quant-ph0511197/Ni.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %\documentclass[11pt]{article}
3: %\documentstyle[11pt,tighten,epsfig,graphicx,color]{article}
4: %\documentclass[11pt]{article}
5: %\usepackage{amssymb,xspace,epsfig,graphicx,color}
6: \documentclass[prd,12pt]{revtex4}
7: \usepackage{amssymb}%,aps,floats,tighten,epsfig,graphicx,color
8: \usepackage{amsmath}
9: \usepackage{graphicx}
10: \usepackage{color}
11: \DeclareGraphicsExtensions{.jpg,.jpeg,.pdf,.png,.mps,.eps,.ps}
12: 
13: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
14: 
15: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
16: %\def\thefootnote{\fnsymbol{footnote}}
17: \newcommand{\beq}{\begin{equation}}
18: \newcommand{\eeq}{\end{equation}}
19: \newcommand{\bea}{\begin{eqnarray}}
20: \newcommand{\ena}{\end{eqnarray}}
21: \newcommand{\etal}{{\it et al.}}
22: \newcommand{\ie}{{\it i.e.}}
23: \newcommand{\eg}{{\it e.g.}}
24: \newcommand{\lsim}{\mathrel{\mathop{\kern 0pt \rlap
25: {\raise.2ex\hbox{$<$}}}
26: \lower.9ex\hbox{\kern-.190em $\sim$}}}
27: \newcommand{\gsim}{\mathrel{\mathop{\kern 0pt \rlap
28: {\raise.2ex\hbox{$>$}}}
29: \lower.9ex\hbox{\kern-.190em $\sim$}}}
30: \newcommand{\un}[1]{_{\mbox{\scriptsize #1}}}
31: \newcommand{\puis}[1]{$^{#1}$}
32: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
33: 
34: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
35: \newcommand{\app}[3]{{\it Astropart.\ Phys.}\ #2{\bf #1} #3 }
36: \newcommand{\hepex}[1]{{\tt hep-ex/#1}}
37: \newcommand{\hepth}[1]{{\tt hep-th/#1}}
38: \newcommand{\hepph}[1]{{\tt hep-ph/#1}}
39: \newcommand{\physics}[1]{{\tt physics/#1}}
40: \newcommand{\astroph}[1]{{\tt astro-ph/#1}}
41: \newcommand{\prep}[3]{#2\ {\it Phys.\ Rep.}\ {\bf #1} #3 }
42: \newcommand{\plb}[3]{#2\ {\it Phys.\ Lett.\ B}\ {\bf #1} #3 }
43: \newcommand{\npb}[3]{#2\ {\it Nucl.\ Phys.\ B}\ {\bf #1} #3 }
44: \newcommand{\cpc}[3]{#2\ {\it Comm.\ Phys.\ Comm.}\ {\bf #1} #3 }
45: \renewcommand{\apj}[3]{#2\ {\it Astrophys.\ J.}\ {\bf #1} #3 }
46: \newcommand{\aeta}[3]{#2\ {\it Astron.\ {\&}\ Astrophys.}\ {\bf #1} #3 }
47: \newcommand{\pr}[3]{#2\ {\it Phys.\ Rev.}\ {\bf #1} #3 }
48: \renewcommand{\prl}[3]{#2\ {\it Phys.\ Rev.\ Lett.} {\bf #1} #3 }
49: \renewcommand{\prd}[3]{#2\ {\it Phys.\ Rev.\ D}\ {\bf #1} #3 }
50: \renewcommand{\pra}[3]{#2\ {\it Phys.\ Rev.\ A}\ {\bf #1} #3 }
51: \renewcommand{\rmp}[3]{#2\ {\it Rev.\ Mod.\ Phys.}\ {\bf #1} #3 }
52: \newcommand{\rnc}[3]{#2\ {\it Riv.\ Nuovo\ Cim.}\ {\bf #1} #3 }
53: \newcommand{\zfpc}[3]{#2\ {\it Z.\ Phys.\ C}\ {\bf #1} #3 }
54: \newcommand{\jpb}[3]{#2\ {\it J.\ Phys.\ B:\ Atom Mol. Opt. Phys.} {\bf #1} #3 }
55: \newcommand{\href}[2]{#1}
56: %\newcommand{\email}[1]{\tt #1}
57: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
58: 
59: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60: \newcommand{\dl}{\mbox{$d_{\rm L}$}}
61: \newcommand{\msol}{\mbox{$M_{\odot}$}}
62: \newcommand{\lsol}{\mbox{$L_{\odot}$}}
63: \newcommand{\pM}{\mbox{$P_{\rm M}$}}
64: \newcommand{\pL}{\mbox{$P_{\Lambda}$}}
65: \newcommand{\rhoM}{\mbox{$\rho_{\rm M}$}}
66: \newcommand{\rhoL}{\mbox{$\rho_{\Lambda}$}}
67: \newcommand{\rclose}{\mbox{$\rho_{C}^{\, 0}$}}
68: \newcommand{\omegaM}{\mbox{$\Omega_{\rm cdm}$}}
69: \newcommand{\omegaB}{\mbox{$\Omega_{\rm b}$}}
70: \newcommand{\omegaL}{\mbox{$\Omega_{\Lambda}$}}
71: \newcommand{\omegaK}{\mbox{$\Omega_{\rm K}$}}
72: \newcommand{\omegachi}{\mbox{$\Omega_{\chi}$}}
73: \newcommand{\Ochi}{\mbox{$\Omega_{\chi} \, h^{2}$}}
74: \newcommand{\etaphi}{\mbox{$\eta_{\Phi}$}}
75: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
76: \newcommand{\yF}{\mbox{$y_{F}$}}
77: \newcommand{\xF}{\mbox{$x_{F}$}}
78: \newcommand{\TF}{\mbox{$T_{F}$}}
79: \newcommand{\Sa}{\mbox{$\sigma_{\rm an}$}}
80: \newcommand{\nA}{\mbox{$n_{\chi}$}}
81: \newcommand{\nAe}{\mbox{$n_{\chi}^{\, 0}$}}
82: \newcommand{\fA}{\mbox{$f_{\chi}$}}
83: \newcommand{\fF}{\mbox{$f_{F}$}}
84: \newcommand{\FA}{\mbox{$F_{\chi}$}}
85: \newcommand{\fAe}{\mbox{$f_{\chi}^{\, 0}$}}
86: \newcommand{\fAasy}{\mbox{$f_{\chi}^{\, \rm asy}$}}
87: \newcommand{\rchi}{\mbox{$\rho_{\chi}^{\, 0}$}}
88: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
89: 
90: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
91: % Some color definitions
92: \definecolor{cyan}{cmyk}{1.,0.,0.,0.5}
93: \definecolor{magenta}{cmyk}{0.,1.,0.,0.5}
94: \definecolor{verdatre}{cmyk}{0.5,0.,0.5,0.5}
95: \definecolor{yellow}{cmyk}{0.,0.,0.2,0.0}
96: \definecolor{rouge}{cmyk}{0.,0.4,0.6,0.0}
97: \definecolor{orange}{cmyk}{0.,0.5,0.5,0.}
98: \definecolor{violet}{rgb}{0.5,0.,0.5}
99: \renewcommand\baselinestretch{1.5}
100: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
101: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
102: 
103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
104: %                      TITLE PAGE + ABSTRACT
105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
106: \begin{document}
107: %\begin{CJK*}{GBK}{song}
108: 
109: \noindent
110: \title{Reduced Dirac Equation And Lamb Shift As An Off-mass-shell \\
111: Effect In Quantum Electrodynamics}
112:  \vskip 1.cm
113: \author{Ni Guang-jiong $^{\rm a,b}$}%£¨Ä߹⾼£©
114: \email{\ pdx01018@pdx.edu}
115: \affiliation{$^{\rm a}$ Department of Physics, Fudan University, Shanghai, 200433, China\\
116: $^{\rm b}$ Department of Physics, Portland State University, Portland, OR97207, U. S. A.}
117: 
118: \author{Xu Jianjun}%£¨Ð콨¾ü£©
119: \email{\ xujj@fudan.edu.cn}
120: \affiliation{Department of Physics, Fudan University, Shanghai, 200433, China}
121: 
122: \author{Lou Senyue$^{\rm c,d}$}%£¨Â¥É­ÔÀ£©
123: \email{\ sylou@sjtu.edu.cn}
124: \affiliation{$^{\rm c}$ Department of Physics, Shanghai Jiao Tong University, Shanghai, 200030, China\\
125: $^{\rm d}$ Department of Physics, Ningbo University, Ningbo 315211, China}
126: 
127: \vskip 0.5cm
128: \date{\today}
129: %\centerline{July 18,2002}
130: 
131: \vskip 0.5cm
132: \begin{abstract}
133: Based on the precision experimental data of energy-level
134: differences in hydrogenlike atoms, especially the $1S-2S$ transition
135: of hydrogen and deuterium, the necessity of introducing a reduced
136: Dirac equation with reduced mass as the substitution of
137: original electron mass is stressed. Based on new cognition about the essence of special relativity, we
138: provide a reasonable argument for reduced Dirac equation to have two symmetries, the
139: invariance under the (newly defined) space-time inversion and that
140: under the pure space inversion, in a noninertial frame. By using reduced
141: Dirac equation and within the framework of quantum electrodynamics in covariant
142: form, the Lamb shift can be evaluated (at one-loop level) as the
143: radiative correction on a bound electron staying in an
144: off-mass-shell state--a new approach eliminating the infrared
145: divergence. Hence the whole calculation, though with limited
146: accuracy, is simplified, getting rid of all divergences and free
147: of ambiguity.\\
148: {\bf Keywords}:\;Reduced Dirac Equation, Lamb shift, off-mass-shell\\
149: {\bf PACC}:\; 0365, 1110G, 1220D
150: \end{abstract}
151: 
152: \maketitle \vskip 1cm
153: %\numberwithin{equation}{section}
154: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
155: % INTRODUCTION
156: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
157: \section{Introduction}
158: \label{sec:introduction}
159: 
160: \vskip 0.1cm As is well known, the Dirac equation for electron in
161: a hydrogenlike atom is usually treated as a one-body equation with
162: the nucleus being an inert core having infinite mass and exerting
163: a potential $V(r)=-\frac{Z\alpha}{r}\ \ (\hbar=c=1)$ on the
164: electron. Then the rigorous solution of energy levels reads[1]:
165: \bea
166:  E_{nj}&=&m_ef(n,j)   \label{1-1}\\
167:  f(n,j)&=&\left[1+\frac{(Z\alpha)^2}{(n-\beta)^2}\right]^{-\frac{1}{2}}\label{1-2}\\
168:  \beta&=&j+\frac{1}{2}-\sqrt{(j+\frac{1}{2})^2-(Z\alpha)^2}
169:  \label{1-3}
170: \ena
171: 
172: where $j$ is the total angular momentum. The expansion of $f(n,j)$ to
173: the power of $(Z\alpha)^6$ is given as \cite{1}
174: 
175: \beq\begin{array}{l}
176: f(n,j)=1-\frac{(Z\alpha)^2}{2n^2}-\frac{(Z\alpha)^4}{2n^3}\left(\frac{1}{j+\frac{1}{2}}
177: -\frac{3}{4n}\right)\\
178: -\frac{(Z\alpha)^6}{8n^3}\left[\frac{1}{(j+\frac{1}{2})^3}
179: +\frac{3}{n(j+\frac{1}{2})^2}+\frac{5}{2n^3}-\frac{6}{n^2(j+\frac{1}{2})}\right]+\cdots
180: \label{1-4}
181: \end{array}\eeq
182: 
183: Obviously, besides the rest energy of the electron given by the
184: first term, the second term has exactly the form of Bohr energy
185: level except that the mass $m_e$ must be replaced by the reduced
186: mass
187: 
188: \beq \mu=\frac{m_em_N}{m_e+m_N}\equiv\frac{m_em_N}{M}
189: \label{1-5}
190: \eeq
191: 
192: with $m_N$ being the mass of the nucleus and $M=m_e+m_N$.
193: 
194: However, as discussed in Refs.\cite{1} and \cite{2}, the concept
195: of reduced mass in relativistic quantum mechanics (RQM) is
196: ambiguous to some extent. Beginning from 1950's, a number of
197: authors have been devoting a great effort at the level of two-body
198: RQM and that of quantum electrodynamics (QED) to take account of the recoil effect   \cite{3,4,1}, incorporating their
199: results in a compact form (to order of $\alpha^4$):
200: 
201: \beq
202: E=M+\mu[f(n,j)-1]-\frac{{\mu}^2}{2M}[f(n,j)-1]^2+\frac{(Z\alpha)^4{\mu}^3}{2n^3m_N^2}
203: \left[\frac{1}{j+\frac{1}{2}}-\frac{1}{l+\frac{1}{2}}\right](1-{\delta}_{l0})
204: \label{1-6}
205: \eeq
206: A comprehensive review on the theory of hydrogenlike atoms can be found in Ref.\cite{27}.
207: The aim of this paper is two-fold: First, based on the
208: experimental data of hydrogen $1S-2S$ transition frequency
209: \cite{5} and its isotope shift of hydrogen and deuterium \cite{6},
210: we stress the necessity of the introduction of reduced mass $\mu$
211: (section II) before we are able to argue the reasonableness of
212: introducing a "reduced Dirac equation" with $\mu$ as the
213: substitution of $m_e$ (section III). Second, based on above
214: conception, we will present a calculation of Lamb Shift (LS) as an
215: off-mass-shell effect by performing the evaluation of self-energy
216: diagrams of electron (section IV) and photon (section V) as well
217: as the vertex function (section VI) at the one-loop level of
218: QED in covariant form. The new insight
219: of our calculation is focused on the regularization renormalization
220: method (RRM). As initiated by J-F Yang \cite{7} and elaborated in
221: a series of papers (\cite{8:a,8:b,9,24,25,26} and references therein), we can get
222: rid of all ultra violet divergences in the calculation of quantum
223: field theory (QFT). Furthermore, in this paper, we will be able to
224: get rid of the annoying infrared divergence in the vertex function
225: by treating the electron moving off its mass-shell to certain
226: extent which is fixed through the evaluation of self-energy
227: diagram or by the Virial theorem. Based on above improvements, the
228: one-loop calculation yields values of LS in a simple but
229: semi-quantitative way (section VII and VIII). Although the
230: accuracy is limited at one-loop level, we hope our approach could
231: be served as a new starting point for calculations at high-loop
232: order to get accurate results at a comparably low labor cost. The
233: final section IX and Appendix will contain a summary and discussion.
234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
235: % Transition
236: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
237: \section{The $1S-2S$ Transition of Atomic Hydrogen and Deuterium}
238: \label{sec:transition}
239: 
240: \vskip 0.1cm
241: 
242: In the last decade, thanks to remarkable advances in high
243: resolution laser spectroscopy and optical frequency metrology, the
244: $1S-2S$ two-photon transition in atomic hydrogen $H$ (or deuterium
245: $D$) with its natural linewidth of only $1.3 Hz$ had been measured
246: to a very high precision. In 1997, Udem \etal determined the
247: $1S-2S$ interval of $H$ being \cite{5}
248: 
249: \beq f^{(H)}(1S-2S)=2466061413187.34(84)\ \ kHz
250: \label{2-1}
251: \eeq
252: 
253: Even four years earlier, Schmidt-Kalar \etal ~measured the
254: isotope-shift of the $1S-2S$ transition of $H$ and $D$ to an
255: accuracy of $3.7\times 10^{-8}$\cite{6}, giving (as quoted in
256: \cite{10}):
257: 
258: \beq \Delta f\equiv f^{(D)}(2S-1S)-f^{(H)}(2S-1S)=670994337(22)\ \ kHz
259: \label{2-2}
260: \eeq
261: (In 1998, Huber \etal measured a more accurate data \cite{28}: $670994334.64(15)\ kHz$).
262: which is of the order of $10^{-4}$ in comparison with Eq.(\ref{2-1}).
263: As pointed out in Ref.\cite{6}, this $671\ GHz$ isotope-shift can
264: be ascribed almost entirely to the different masses of proton
265: ($p$) and deuteron ($d$). And the nuclear volume effects become
266: important because the QED effects cancel considerably in the
267: isotope shift.
268: 
269: Here, we wish to emphasize that in the first approximation, both
270: experimental data (\ref{2-1}) and (\ref{2-2}) can be well
271: accounted for by simply resorting to Eq.(\ref{1-1}) with $m_e$
272: replaced by the reduced mass
273: 
274: \beq \mu_H=\frac{m_em_p}{m_e+m_p},\ \ \
275: \mu_D=\frac{m_em_d}{m_e+m_d}
276: \label{2-3}
277: \eeq
278: 
279: for $H$ and $D$ respectively.
280: 
281: Indeed, adopting the following updated values \cite{10,11,12,13}
282: 
283: \bea
284: \alpha&=&(137.03599944)^{-1},\ \ \
285: {\alpha}^2=0.532513542\times10^{-4}\\
286: {\alpha}^4&=&0.283570673\times10^{-8},\ \ \
287: {\alpha}^6=0.151005223\times10^{-12}\\
288: m_e&=&0.51099906\ \ MeV=1.2355897\times10^{20}\ \ Hz\\
289: R_{\infty}&=&\frac{1}{2}{\alpha}^2m_e=3.28984124\times10^{15}\ \
290: Hz\\
291: \frac{m_p}{m_e}&=&1836.1526665
292:  \label{data}
293:  \ena
294: 
295: and denoting
296: \bea
297: \frac{m_e}{m_p}&=&b_H=5.446170255\times10^{-4},\ \ \
298: \frac{1}{1+b_H}=0.999455679\\
299: \frac{m_e}{m_d}&=&b_D=2.724436319\times10^{-4},\ \ \
300: \frac{1}{1+b_D}=0.99972763
301: \label{data2}
302: \ena
303: 
304: we can calculate the energy difference of $2S$ and $1S$ of $H$
305: through Eq.(\ref{1-1}) with $m_e$ replaced by $\mu_H$ (the superscript
306: RDE refers to the reduced Dirac equation)
307: 
308: \bea
309: \Delta E^{RDE}_{H}(2S-1S)&=&\mu_H[f(2,1/2)-f(1,1/2)]\nonumber \\
310: &=&\frac{m_e}{1+b_H}(1.996950159\times10^{-5})\nonumber \\
311: &=&1.2355897\times10^{20}\times0.999455679\times1.996950159\times10^{-5}\nonumber \\
312: &=&2.466067984\times10^{15}\ \ Hz
313: \label{2-4}
314: \ena
315: 
316: which is only a bit larger than the experimental data Eq.(\ref{2-1})
317: with accuracy $3\times10^{-6}$. However, a more stringent test of
318: RDE should be the isotope shift of $H$ and $D$. We have
319: 
320: \beq
321: \frac{1}{1+b_D}-\frac{1}{1+b_H}=(b_H-b_D)-(b^2_H-b^2_D)+(b^3_H-b^3_D)+\cdots=2.719511528\times10^{-4}
322: \label{2-5}
323: \eeq
324: 
325: \beq
326: \Delta
327: E^{RDE}_{D-H}=(\mu_D-\mu_H)[f(2,1/2)-f(1,1/2)]=6.7101527879\times10^{11}\
328: \ Hz
329: \label{2-6}
330: \eeq
331: 
332: which has only a discrepancy larger than the experimental data,
333: Eq.(\ref{2-2}) by $20.941\ MHz$ with accuracy $3\times10^{-5}$. Of
334: course, it is still not satisfied in an analysis of high precision
335: \cite{6}. Let us resort to the Eq.(6), where the third term does
336: provide a further modification:
337: 
338: \bea
339: &-&\frac{1}{2}m_e[\frac{b_D}{(1+b_D)^3}-\frac{b_H}{(1+b_H)^3}]\{[f(2,1/2)-1]^2-[f(1,1/2)-1]^2\}\nonumber\\
340: &=&\frac{1}{2}m_e[(b_H-b_D)-3(b^2_H-b^2_D)+\cdots](-6.646361554\times10^{-10})=-11.176
341: \ MHz
342: \label{2-7}
343: \ena
344: 
345: which brings the discrepancy between the theory and experimental
346: down to less than $10\ MHz$.
347: 
348: Although the detail explanation for this discrepancy remains quite
349: complicated\cite{6}, the above comparison is enough to convince us
350: that the inevitable appearance of reduced mass in the RDE or
351: Eq.(6) is by no means a simple fortune. It must have a deep reason
352: from a theoretical point of view. Notice further that once the
353: conditions $m_e\ll m_p$ and $m_e\ll m_d$ hold, the difference  of
354: spin between $p$ and $d$ seems not so important. So in next
355: section, we will strive to justify the reduced Dirac equation on a
356: reasonable basis. Of course, it is still an approximate one, but
357: seems much better than the original Dirac equation when dealing
358: with hydrogenlike atoms.
359: 
360: 
361: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
362: % Reduced Dirac Equation
363: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
364: \section{Reduced Mass and Reduced Dirac Equation}
365: \label{sec:decoupling}
366: 
367: Consider a system of two particles with rest masses $m_1$ and
368: $m_2$. Their coordinates in the center-of-mass (CM) system are ${\mathbf r}_1$ and ${\mathbf r}_2$ respectively,
369: as shown in Fig.1. If there is a potential $V(r)=V(|{\mathbf
370: r_1-r_2}|)$ between them, two equations $m_1\ddot{\mathbf
371: r}_1=-\nabla_rV(r)$ and $m_2\ddot{\mathbf r}_2=\nabla_rV(r)$ will
372: reduce to one:
373: 
374: \beq
375: \mu\frac{d^2{\mathbf r}}{dt^2}=-\nabla_rV(r), \ \ \
376: (\mu=\frac{m_1m_2}{m_1+m_2})
377: \label{3-1}
378: \eeq
379: 
380: At first sight, the definition of center-of -mass (CM) in
381: classical mechanics $m_1r_1=m_2r_2$ becomes doubtful in the theory
382: of special relativity (SR) because the mass is no longer a
383: constant. But actually, we can still introduce the coordinate of
384: CM in the laboratory coordinate system (LCS) (with ${\mathbf
385: r}'_1$ and ${\mathbf r}'_2$ being the coordinates of $m_1$ and
386: $m_2$):
387: 
388: \beq
389: R=\frac{1}{M}(m_1{\mathbf r}'_1+m_2{\mathbf r}'_2)=(X,Y,Z),\
390: \ \ (M=m_1+m_2)
391: \label{3-2}
392: \eeq
393: 
394: and the relative coordinate of $m_1$ and $m_2$ (${\mathbf r}_i={\mathbf r}'_i-{\mathbf R}, \, i=1,2$) :
395: 
396: \beq
397: {\mathbf r}={\mathbf r}'_1-{\mathbf r}'_2={\mathbf
398: r}_1-{\mathbf r}_2=(x,y,z)
399: \label{3-3}
400: \eeq
401: 
402: Here the motion of CM in the LCS is assumed to be slow and so
403: 
404: \beq
405: \frac{\partial}{\partial
406: x'_1}=\frac{m_1}{M}\frac{\partial}{\partial
407: X}+\frac{\partial}{\partial x},\ \ \ \frac{\partial}{\partial
408: x'_2}=\frac{m_2}{M}\frac{\partial}{\partial
409: X}-\frac{\partial}{\partial x}
410: \label{3-4}
411: \eeq
412: 
413: Notice that the momentum $\mathbf P$ of CM and the relative
414: momentum ${\mathbf p}_r$ becomes operator in quantum mechanics
415: (QM) without explicit dependence on mass:
416: 
417: \beq
418: {\mathbf P}=-i\hbar\nabla_{\mathbf R},\ \ \ {\mathbf
419: p}_r=-i\hbar\nabla_{\mathbf r}
420: \label{3-5}
421: \eeq
422: 
423: Thus the momenta of $m_1$ and $m_2$ in laboratory coordinate
424: system (LCS) read:
425: 
426: \beq
427: {\mathbf p}'_1=-i\hbar\nabla_{{\mathbf
428: r}'_1}=\frac{m_1}{M}{\mathbf P}+{\mathbf p}_r,\ \ \ {\mathbf
429: p}'_2=-i\hbar\nabla_{{\mathbf r}'_2}=\frac{m_2}{M}{\mathbf
430: P}-{\mathbf p}_r
431: \label{3-6}
432: \eeq
433: 
434: Since the center-of-mass coordinate system (CMCS) is also an inertial frame which can be transformed
435: from the LCS via a linear Lorentz transformation, it is defined by
436: the condition that $\mathbf P=0$ in CMCS. In other words, CMCS is
437: defined by the condition ${\mathbf p}_1+{\mathbf p}_2=0$, or from
438: Eq.(\ref{3-6}):
439: 
440: \beq
441: {\mathbf p}'_1=-i\hbar\nabla_{{\mathbf r}'_1}={\mathbf p}_r,\
442: \ \ {\mathbf p}'_2=-i\hbar\nabla_{{\mathbf r}'_2}=-{\mathbf p}_r
443: \label{3-7}
444: \eeq
445: 
446: Evidently, the above definition of CMCS remains valid in the realm
447: of relativistic QM (RQM) even the exact meaning of CM seems
448: obscure to some extent due to the conjugation relation of $a$ particle's position and its momentum, see Fig.1.
449: 
450: Now, from Eq.(\ref{3-7}), it is natural to replace $\mathbf p_1$ and
451: $\mathbf p_2$ by ${\mathbf p}_r$, reducing the two-particle
452: degrees of freedom to one. In the meantime, the origin of CMCS is
453: discarded, it is substituted  by the position of $m_2$ (${\mathbf
454: r}={\mathbf r}_1-{\mathbf r}_2$). We will call the system
455: associated with $\mathbf r$ the relative motion coordinate system
456: (RMCS), which should be viewed as a deformation of CMCS. The
457: transformation from CMCS to RMCS is by no means a linear one.
458: Rather, the origin of RMCS ($m_2$) is moving non-uniformly in the CMCS.
459: Therefore, while rest masses $m_1$ and $m_2$ remain the same in
460: both LCS and CMCS, they reduce to one mass
461: $\mu=\frac{m_1m_2}{m_1+m_2}$ for $m_1$ in RMCS (or for $m_2$ if
462: $m_1$ is chosen as the origin of RMCS).
463: 
464: Let us express the total energy $E=E_1+E_2$ in CMCS in terms of
465: $p_r$ and reduced mass $\mu$ ($\mu=\frac{m_1m_2}{M},\
466: M=m_1+m_2$), where
467: 
468: \beq
469: E_1=\sqrt{m_1^2+p_1^2}=\sqrt{m_1^2+p_r^2},\ \ \
470: E_2=\sqrt{m_2^2+p_2^2}=\sqrt{m_2^2+p_r^2}
471: \label{3-8}
472: \eeq
473: 
474: Treating all $p_1,p_2$ and $p_r$ being $c$-numbers, we have
475: 
476: \beq
477: E^2=(E_1+E_2)^2=M^2+\frac{M}{\mu}p_r^2+\frac{1}{4\mu^2}p_r^4(4-\frac{M}{\mu})+\cdots
478: \label{3-9}
479: \eeq
480: where the expansion in $p_r$ is kept to the order
481: of $p_r^4$. Two extreme cases will be considered separately:
482: 
483: {\bf A}. $m_2\gg m_1,\ \  \mu\lesssim m_1, \ \ M\gg\mu$: \bea
484: E^2&=&M^2\left[1+\frac{1}{\mu
485: M}p_r^2-\frac{1}{4M\mu^3}p_r^4(1-\frac{4\mu}{M})+\cdots\right]\nonumber\\
486: E&=&M\left[1+\frac{1}{2\mu
487: M}p_r^2-\frac{1}{8M\mu^3}p_r^4(1-\frac{3\mu}{M})+\cdots\right]\nonumber\\
488: &=&M+\frac{1}{2\mu}p_r^2-\frac{1}{8\mu^3}p_r^4+\cdots\nonumber\\
489: &\simeq &M-\mu+\sqrt{\mu^2+p_r^2}\simeq m_2+(m_1-\mu)+\sqrt{\mu^2+p_r^2}
490: \label{3-10}
491: \ena
492: 
493: \beq
494: E'\equiv E-m_2=(m_1-\mu)+\sqrt{\mu^2+p_r^2}
495: \label{3-11}
496: \eeq
497: 
498: {\bf B}. $m_1=m_2=m,\ \ \mu=\frac{m}{2},\ \  M=2m=4\mu$
499: Then to
500: the accuracy of $p_r^4$, we have :
501: 
502: \bea
503: E^2&=&M^2+\frac{M}{\mu}p_r^2=4m^2+4p_r^2\nonumber\\
504: E&=&2m+\frac{1}{2\mu}p_r^2-\frac{1}{32\mu^3}p_r^4+\cdots\nonumber\\
505: E'&\equiv &E-M=\frac{1}{2\mu}p_r^2-\frac{1}{32\mu^3}p_r^4\simeq
506: \frac{1}{2\mu}p_r^2,\ \ \ (\mbox{if}\ \ p_r^2\ll \mu^2)
507: \label{3-12}
508: \ena
509: 
510: It is interesting to see that after introducing $\mu$ and $p_r$,
511: the energy $E'$ in RMCS looks  quite "relativistic" in the case A
512: whereas it looks rather "non-relativistic" in the case B even both
513: of them are derived from the relativistic expressions, Eq.(\ref{3-8}),
514: approximately.
515: 
516: Since the RMCS is not an inertial system, the original mass of
517: $m_1$ in CM changes abruptly to $\mu$ as shown in Eq.(\ref{3-11}).
518: How can we derive the reduced Dirac equation (RDE) in RMCS?
519: Fortunately, we already found a basic symmetry, the space-time
520: inversion symmetry, which not only serves as the essence of
521: special relativity (SR), but also goes beyond it to derive the original Dirac equation and the
522: tachyon theory for neutrinos \cite{14,15,16,17}. Based on this symmetry,
523: we are going to derive the equation in RQM for case either A or B
524: respectively.
525: 
526: Let us consider case B ($m_1\simeq m_2$) first. The motivation is
527: stemming from the success of using the Schr\"{o}dinger equation to
528: heavy-quarkoniums like $c\bar{c}$ and $b\bar{b}$ in particle
529: physics (\cite{18}, see also \cite{15} \ \S 9.5 D). Ignoring the
530: spin of both $m_1$ and $m_2$, we assume the coupling equations
531: in laboratory system for the two-particle system as:
532: 
533: \beq
534: \left\{
535: \begin{array}{ll}
536: i\hbar\frac{\partial\varphi}{\partial
537: t}&=(m_1+m_2)c^2\varphi+V(|{\mathbf
538: r'_1-\mathbf r'_2}|)(\varphi+\chi)-(\frac{\hbar^2}{2m_1}\nabla^2_{\mathbf
539: r'_1}+\frac{\hbar^2}{2m_2}\nabla^2_{\mathbf r'_2})(\varphi+\chi)\\
540: i\hbar\frac{\partial\chi}{\partial
541: t}&=-(m_1+m_2)c^2\chi-V(|{\mathbf
542: r'_1-\mathbf r'_2}|)(\varphi+\chi)+(\frac{\hbar^2}{2m_1}\nabla^2_{\mathbf
543: r'_1}+\frac{\hbar^2}{2m_2}\nabla^2_{\mathbf r'_2})(\varphi+\chi)
544: \end{array}
545: \right.
546: \label{3-13}
547: \eeq
548: 
549: where $\varphi=\varphi({\mathbf r'_1,\mathbf r'_2},t)$ and
550: $\chi=\chi({\mathbf r'_1,\mathbf r'_2},t)$ are hidden "particle" and
551: "antiparticle" fields of the two-particle system (From now on, the ${\mathbf r}'_i(i=1,2)$ is the flowing coordinate of "fields"
552: in QM, \ie, that of "fictitious point particles". See Fig.1).
553: Eq.(\ref{3-13}) remains invariant under the (newly defined) space-time inversion
554: (${\mathbf r'_1\rightarrow-\mathbf r'_1,
555: \mathbf r'_2\rightarrow-\mathbf r'_2},t\rightarrow-t$):
556: 
557: \beq \left\{
558: \begin{array}{ll}
559: \varphi({-\mathbf r'_1,-\mathbf r'_2},-t)&\longrightarrow\chi({\mathbf
560: r'_1,\mathbf r'_2},t)\\
561: \chi({-\mathbf r'_1,-\mathbf r'_2},-t)&\longrightarrow\varphi({\mathbf
562: r'_1,\mathbf r'_2},t)
563: \end{array}\right.
564: \label{3-14}
565: \eeq
566: 
567: \beq
568: V({-\mathbf r'_1,-\mathbf r'_2},-t)\longrightarrow V({\mathbf r'_1,\mathbf r'_2},t)
569: \label{3-15}
570: \eeq
571: 
572: Note that, however, the time $t$ is not contained in $V$ explicitly. Eq.(\ref{3-15}) merely means that both $m_1$ and $m_2\ (m_1\approx m_2)$
573: transform into their antiparticles under the space-time inversion.
574: Actually, the hidden antiparticle field $\chi$ enhances in nearly
575: equal strength in $m_1$ and $m_2$ when fictitious particles' velocities
576: increase with the enhancement of attractive potential $V(r)$.
577: 
578: After introducing the CM coordinate ${\mathbf
579: R}=\frac{1}{M}(m_1{\mathbf r'}_1+m_2{\mathbf r'}_2)$, ($M=m_1+m_2$)
580: and relative coordinate $\mathbf r=\mathbf r'_1-\mathbf r'_2$, and setting
581: 
582: \beq
583: \varphi=\Phi+i\frac{\hbar}{Mc^2}\dot{\Phi},\ \ \ \
584: \chi=\Phi-i\frac{\hbar}{Mc^2}\dot{\Phi}
585: \label{3-16}
586: \eeq
587: 
588: we find ($\mu=\frac{m_1m_2}{M}$)
589: 
590: \beq
591: \ddot{\Phi}-c^2\nabla^2_R\Phi-c^2\frac{M}{\mu}\nabla^2_r\Phi+\frac{1}{\hbar^2}(M^2c^4+2VMc^2)\Phi=0
592: \label{3-17} \eeq
593: 
594: Its stationary solution reads
595: 
596: \beq
597: \Phi({\mathbf R,\mathbf r},t)=\psi(\mathbf
598: r)\exp\left[\frac{i}{\hbar}({\mathbf P\cdot \mathbf R}-Et)\right]
599: \label{3-18}
600: \eeq
601: 
602: where $E$ is the total energy of the system while $\mathbf P$ the
603: momentum of CM. The reduced "one-body" equation for $\psi(\mathbf r)$ turns out to
604: be: \footnotemark[1]
605: \footnotetext[1]{With Eq.(\ref{3-15}), Eq.(\ref{3-17}) is invariant under the space-time inversion (${\bf r}\to -{\bf r},
606: t\to -t$). Equivalently, under the mass inversion ($m_1\to -m_1,m_2\to -m_2$), Eq.(\ref{3-17}) and Eq.(\ref{3-19}) remain
607: invariant in the sense that not only $\mu\to -\mu, M\to -M$, but also $V({\bf r})\to -V({\bf r}), \varepsilon\to -\varepsilon$.
608: Notice that, however, the simultaneous inversion of $m_1$ and $m_2$ implies $m_1\simeq m_2$, so both particles change under their
609: mutual interaction $V({\bf r})$ simultaneously. Here $V$, being the "internal potential energy" of two-body system, was called
610: as a "scalar potential". We see that either the invariance under the space-time inversion or that under the mass inversion
611: is capable of showing the particle-antiparticle symmetry (\ie, relativistic nature) of a system essentially. }
612: 
613: \beq
614: \left\{
615: \begin{array}{ll}
616: &\left[-\dfrac{\hbar^2}{2\mu}\nabla^2_{\mathbf r}+V(\mathbf r)\right]\psi(\mathbf
617: r)=\varepsilon\psi(\mathbf r)\\[4mm]
618: &\varepsilon=\dfrac{1}{2Mc^2}(E^2-M^2c^4-{\mathbf P}^2c^2)
619: \end{array}
620: \right.
621: \label{3-19}
622: \eeq
623: 
624: We set $\mathbf P=0$ (\ie  turn to CMCS) and denote the binding
625: energy $B=Mc^2-E$, yielding:
626: 
627: \beq
628: B=Mc^2\left[1-(1+\frac{2\varepsilon}{Mc^2})^{1/2}\right]=-\varepsilon+\frac{1}{2}\frac{\varepsilon^2}{Mc^2}-\cdots
629: \label{3-20}
630: \eeq
631: 
632: Notice that although Eq.(\ref{3-19}) looks like a
633: "non-relativistic" stationary Schr\"{o}dinger equation, it is
634: essentially relativistic. This can be seen from its remarkable
635: property that the eigenvalue $\varepsilon$ has a lower bound
636: $-\frac{1}{2}Mc^2$, corresponding to $E_{\mbox{min}}=0\
637: (B_{\mbox{max}}=M)$!
638: 
639: An example is: consider "positronium" composed of $e^+$ and $e^-$ with charge $Ze$ and $-Ze$ respectively. Once when the "fictitious
640: charge number" $Z$ increases from $1$ to $Z_{max}=(\frac{4}{\alpha^2})^{1/4}=16.555$, the whole bound system would have lowest ground
641: energy $E_{min}=0$!
642: So Eq.(\ref{3-19}) is really a relativistic QM equation capable of
643: giving a nonperturbative solution under the strong coupling.
644: 
645: Eq.(\ref{3-19}) provides a justification (realization) of
646: conjecture Eq.(\ref{3-12}) relevant to case B ($m_1\simeq m_2$) where
647: the spin of both particles is merely of second importance.
648: 
649: Now let us turn to case A where $m_2\gg m_1$, taking the spin of
650: $m_1$ into account but ignoring that of $m_2$ as before. Based on
651: the experience in case B, also because of great difficulty to
652: derive the equation starting from the laboratory system for this
653: case A, we directly introduce the reduced Dirac equation (RDE) in
654: the RMCS as a pair of coupled equations of two-component spinors
655: $\varphi({\mathbf r},t)$ and $\chi({\mathbf r},t)$, ($c=\hbar=1$)
656: 
657: \beq
658: \left\{
659: \begin{array}{ll}
660: i\dot{\varphi}&=i{\mathbf \sigma}_1\cdot\nabla_{\mathbf r}\chi+\mu\varphi+V({\mathbf r})\varphi\\
661: i\dot{\chi}&=i{\mathbf \sigma}_1\cdot\nabla_{\mathbf r}\varphi-\mu\chi+V({\mathbf r})\chi
662: \end{array}
663: \right.
664: \label{3-21}
665: \eeq
666: with $\mu$ replacing $m_1$. Here $\mathbf \sigma_1$ are Pauli matrices acting on the spin
667: space of particle $m_1$. Eq.(\ref{3-21}) is invariant under the
668: space-time inversion $({\mathbf r\rightarrow-\mathbf r},t\rightarrow-t), \
669: \varphi(-{\mathbf r},-t)\rightarrow \chi({\mathbf r},t),\ \chi(-{\mathbf r},-t)\rightarrow \varphi({\mathbf r},t)$
670: whereas we assume
671: 
672: \beq
673: V(-{\mathbf r},-t)\longrightarrow-V({\mathbf r},t)
674: \label{3-22}
675: \eeq
676: here in contrast to Eq.(\ref{3-15}) for the case B. \footnotemark[2]
677: \footnotetext[2]{For a hydrogenlike atom, $V(r)=-\frac{Ze^2}{r}$ does not contain time $t$ explicitly.
678: Eq.(\ref{3-22}) merely means that under the space-time inversion, the electron transforms into a position whereas the nucleus
679: remains unchanged. See point (a) of section IX. Previously, the $V$ in Eq.(\ref{3-22}) was called as a "vector potential",
680: meaning the "potential energy" of the electron in an "external field" of nucleus. Note that, formally, Eq.(\ref{3-21}) remains
681: invariant under a mass inversion as $\mu\to -\mu, \phi\to\chi, \chi\to\phi$ ($V({\bf r})$ remains unchanged) in a noninertial
682: frame $RMCS$. Actually, since $m_1=m_e\to -m_e$, but $m_2=m_N\to m_N, \mu\to -\mu(1+\frac{2m_e}{M})$. So Eq.(\ref{3-21})
683: has an inaccuracy up to $\frac{2m_e}{M}$ ($<1.1\times 10^{-3}$ for $H$).}
684: 
685: The reasons are as follows: (a) Eq.(\ref{3-21}) should degenerate into the
686: original Dirac equation when $m_2\rightarrow\infty$,
687: $\mu=\frac{m_1m_2}{m_1+m_2}\rightarrow m_1$. (b) Since now $m_2\gg
688: m_1$ (but $m_2\neq\infty$), $m_1$ is moving much faster than $m_2$
689: in the CMCS. Hence the antiparticle field $\chi$ enhances much
690: appreciably in $m_1$ than that in $m_2$, a situation totally
691: different from that in the case B where $m_1\approx m_2$. (c) If
692: instead of Eq.(\ref{3-22}), we still assume $V(-{\mathbf
693: r},-t)\longrightarrow V({\mathbf r},t)$ like Eq.(\ref{3-15}) and
694: change the sign before $V(\mathbf r)$ in the second equation of
695: Eq.(\ref{3-21}) to keep its invariance under the space-time
696: inversion, then we would get an equation which would lead to a
697: reversed fine-structure of atom (\eg, the $P_{1/2}$ state would
698: lie above the $P_{3/2}$ state), a wrong prediction obviously
699: excluded by experiments.
700: 
701: However, one kind of invariance is not enough to fix an equation. Indeed,
702: the beauty of Dirac equation or RDE is hidden in two symmetries:
703: besides the symmetry of space-time inversion, it has another
704: left-right (parity) symmetry. To see it, we define
705: 
706: \beq
707: \xi=\frac{1}{\sqrt{2}}(\varphi+\chi),\ \ \
708: \eta=\frac{1}{\sqrt{2}}(\varphi-\chi)
709: \label{3-23}
710: \eeq
711: 
712: and recast Eq.(\ref{3-21}) into:
713: 
714: \beq
715: \left\{
716: \begin{array}{ll}
717: i\dot{\xi}&=i{\mathbf \sigma}_1\cdot\nabla_{\mathbf r}\xi+\mu\eta+V(\mathbf r)\xi\\
718: i\dot{\eta}&=-i{\mathbf \sigma}_1\cdot\nabla_{\mathbf r}\eta+\mu\xi+V(\mathbf r)\eta
719: \end{array}
720: \right.
721: \label{3-24}
722: \eeq
723: 
724: which is invariant under a pure space inversion (${\mathbf
725: r\rightarrow-\mathbf r},t\rightarrow t$) if assuming
726: 
727: \beq
728: \xi({-\mathbf r},t)\rightarrow\eta({\mathbf r},t),\ \
729: \eta({-\mathbf r},t)\rightarrow\xi({\mathbf r},t),\ \ V(-{\mathbf
730: r})\rightarrow V({\mathbf r})=V(r)
731: \label{3-25}
732: \eeq
733: 
734: The parity invariance of Dirac equation or RDE has a far-reaching
735: consequence that the Dirac particle is always a subluminal
736: one. By contrast, once the parity is violated to maximum, a
737: superluminal particle (tachyon) will emerge. Interestingly enough,
738: any theory capable of treating particle and antiparticle on an equal
739: footing must respect to the common basic symmetry---the
740: invariance of space-time inversion. The new insight of this
741: section is this symmetry can be applied even in a noninertial
742: frame---the RMCS. Of course, the validity of RDE can only be
743: verified by experiments as discussed in section II, although it is
744: still an approximate description of nature like any other theory
745: in physics. For further discussion, see section IX.
746: 
747: \section{Self-Energy Correction of a Bound Electron in Atom}
748: \label{sec:selfenergy}
749: 
750: In our understanding, one important reason why the calculations of
751: QED for electron in a hydrogenlike atom is so complicated lies in
752: the fact that while calculations are performed in the CMCS, the
753: center of potential (the nucleus with mass $m_2=m_N$) undergoes a
754: complex motion. So the recoil effect interwinds with the high-loop
755: correction of QED, as discussed in many chapters of the books
756: \cite{1} and \cite{2}. We will try to find an alternative approach
757: by adopting the RDE and doing calculation in the RMCS. Let us
758: begin with the Feynman diagram integral (FDI) of electron
759: self-energy at one-loop level, adopting the Bjorken-Drell metric
760: and rationalized Gaussian units with electron charge $-e(e>0)$,
761: see Fig.2(a) (\cite{8:a}).
762: 
763: \beq
764: -i\Sigma(p)=(ie)^2\int\frac{d^4k}{(2\pi)^4}\frac{g_{\mu\nu}}{ik^2}
765: \gamma^{\mu}\frac{i}{\not\!\!p-\not\!\!k-\mu}\gamma^{\nu}
766: \label{4-1}
767: \eeq
768: 
769: Here a free electron with reduced mass $\mu$ is moving at a
770: four-dimensional momentum $p$, whose spatial component is just the
771: relative momentum $\mathbf p_r$ discussed in the previous section,
772: $k$ is the momentum of virtual photon. As usual, a Feynman parameter $x$ will bring
773: Eq.(\ref{4-1}) into
774: 
775: \beq
776: -i\Sigma(p)=-e^2\int\frac{d^4k}{(2\pi)^4}\frac{N}{D}
777: \label{4-2}
778: \eeq
779: 
780: \beq
781: \frac{1}{D}=\int^1_0\frac{dx}{[k^2-2p\cdot
782: kx+(p^2-\mu^2)x]^2},\ \ \ N=-2(\not\!\!p-\not\!\!k)+4\mu
783: \label{4-3}
784: \eeq
785: 
786: ($\not\!\!p=p^\mu\gamma_\mu,\ p\cdot k=p^\mu k_\mu$). A shift in
787: momentum integration:$k\rightarrow K=k-xp$ recast Eq.(\ref{4-2})
788: into
789: 
790: \beq
791: -i\Sigma(p)=-e^2\int^1_0dx[-2(1-x)\not\!\!p+4\mu]I
792: \label{4-4}
793: \eeq
794: 
795: with a logarithmically divergent integral (in Minkowski momentum
796: space):
797: 
798: \beq I=\int\frac{d^4K}{(2\pi)^4}\frac{1}{[K^2-M^2]^2},\ \ \
799: M^2=p^2x^2+(\mu^2-p^2)x \label{4-5} \eeq
800: 
801: Our new regularization-renormalization method (RRM) is based on a
802: cognition that the virtual process in the self-energy diagram does provide a radiative correction to the
803: electron mass but only when the electron is off the mass shell,
804: \ie, $p^2\not=\mu^2$. When it is on the mass shell, $p^2=\mu^2$,
805: the appearance of a divergent integral like $I$ in Eq.(\ref{4-5}) is
806: essentially a warning on the fact that to calculate the mass of
807: electron is beyond the ability of perturbative QED.
808: 
809: Let us consider the converse: if $\Sigma(p)$ does modify the
810: electron mass $\mu$ to some extent, it must comes from the
811: divergent integral $I$. However, the latter is a dimensionless
812: number, we can change the unit of $M$ (and $k$) at our disposal
813: without any change in the value of $I$. So any real change of
814: $\mu$ (on the mass shell) is incredible. The deeper reason lies in
815: a "principle of relativity" in epistemology: everything is moving
816: and becomes recognizable only in relationship with other things. What we can understand is either no
817: mass scale or two mass scales, but never one mass scale.
818: For instance, in the famous Gross-Neveu model \cite{19}, a massive
819: fermion is created only in accompanying with the change (phase
820: transition) of its environment (vacuum) which provides another
821: mass scale (a standard weight). Another example is just the change
822: of electron mass from $m_e$ to $\mu_H$ in a hydrogen atom due to
823: the coexistence of atom nucleus---the proton, this change is also
824: a nonperturbative effect.
825: 
826: Therefore, we expected too much in the past. There is no way to evaluate Eq.(\ref{4-5}) unambiguously
827: or pick out some finite and fixed modification on the mass $\mu$.
828: What we can do is to separate the valuable information carried by
829: Eq.(\ref{4-5}) from an arbitrary constant which will be introduced by a simple trick and then
830: fixed by the experimental data of $\mu$. We will see the
831: information telling us exactly how the value of $I$ changes when
832: the electron is moving off the mass shell.
833: 
834: To handle Eq.(\ref{4-5}), we perform a differentiation with respect
835: to the mass-square parameter $M^2$, then the integration with
836: respect to $K$ becomes convergent, yielding:
837: 
838: \beq
839: \frac{\partial I}{\partial
840: M^2}=\frac{-i}{(4\pi)^2}\frac{1}{M^2}
841: \label{4-6}
842: \eeq
843: 
844: which tells us that while the exact value of $I$ remains obscure,
845: its change linked with $M^2$ has a definite meaning. So we
846: reintegrate Eq.(\ref{4-6}) with respect to $M^2$ and arrive at
847: 
848: \beq I=\frac{-i}{(4\pi)^2}(\ln
849: M^2+C_1)=\frac{-i}{(4\pi)^2}\ln\frac{M^2}{\mu_2^2} \label{4-7}
850: \eeq
851: 
852: where an arbitrary constant $C_1=-\ln\mu_2^2$ is introduced
853: ($\mu_2$ should not be confused with the reduced mass $\mu$).
854: Further integration with respect to Feynman parameter $x$ leads to
855: 
856: \beq
857: \begin{array}{ll}
858:  \Sigma(p)&=A+B\not\!\!p\\
859: A&=\frac{\alpha}{\pi}\mu\left[2-2\ln\frac{\mu}{\mu_2}+\frac{(\mu^2-p^2)}{p^2}\ln\frac{(\mu^2-p^2)}{\mu^2}\right]\\
860: B&=\frac{\alpha}{4\pi}\left[2\ln\frac{\mu}{\mu_2}-3-\frac{(\mu^2-p^2)}{p^2}
861: \left[1+\frac{(\mu^2+p^2)}{p^2}\ln\frac{(\mu^2-p^2)}{\mu^2}\right]\right]
862: \end{array}
863: \label{4-8}
864: \eeq
865: 
866: Using the chain approximation, we can derive the modification of
867: electron propagator as
868: 
869: \beq
870: \frac{i}{\not\!\!p-\mu}\rightarrow\frac{i}{\not\!\!p-\mu}\frac{1}{1-\frac{\Sigma(p)}{\not\!\!p-\mu}}
871: =\frac{iZ_2}{\not\!\!p-\mu_R}
872: \label{4-9}
873: \eeq
874: 
875: where
876: 
877: \beq
878: Z_2=\frac{1}{1-B}
879: \label{4-10}
880: \eeq
881: 
882: is the renormalization factor for wave function of electron and
883: 
884: \beq
885: \mu_R=\frac{\mu+A}{1-B}
886: \label{4-11}
887: \eeq
888: 
889: is the renormalized mass of $\mu$. The increment of mass reads
890: 
891: \beq
892: \delta\mu=\mu_R-\mu=\frac{A+\mu B}{1-B}
893: \label{4-12}
894: \eeq
895: 
896: For a free electron (in the atom), the mass-shell condition
897: $p^2=\mu^2$ should lead to
898: 
899: \beq
900: \delta\mu|_{p^2=\mu^2}=\frac{\alpha\mu}{4\pi}(5-6\ln\frac{\mu}{\mu_2})=0
901: \label{4-13}
902: \eeq
903: 
904: as discussed above\footnotemark[3]. So we must set $\mu_2=\mu e^{-5/6}$ which in
905: turn fixes
906: \footnotetext[3]{We will keep the same mass symbol $\mu$ through out high-loop calculations of QED and reconfirm (renormalize)
907: it at every step by experiment. Just like one has to reconfirm his plane ticket before his departure from the airport,
908: he must use the same name through out his entire journey \cite{8:b}.}
909: \beq
910: Z_2|_{p^2=\mu^2}=\frac{1}{1+\frac{\alpha}{3\pi}}\approx
911: 1-\frac{\alpha}{3\pi}
912: \label{4-14}
913: \eeq
914: 
915: However, the above evaluation further provides us with important
916: knowledge of $\delta\mu$ when electron is moving off the
917: mass-shell. Consider the similar diagram in Fig.(2b), we can
918: set on an average meaning that
919: 
920: \beq
921: p^2=\mu^2(1-\zeta)
922: \label{4-15}
923: \eeq
924: 
925: with $\zeta>0$, which implies from Eq.(\ref{4-12}) with Eq.(\ref{4-8})
926: that \cite{20}:
927: 
928: \beq
929: \delta\mu=\frac{\alpha\mu}{4\pi}\frac{(-\zeta+2\zeta\ln\zeta)}{1+\alpha/3\pi}
930: \label{4-16}
931: \eeq
932: 
933: where some terms of the order of $\zeta^2$ or $\zeta^2\ln\zeta$ are
934: neglected since $\zeta\ll1$. Eq.(\ref{4-16}) establishes the
935: correspondence between the mass modification $\delta\mu$ and the
936: parameter $\zeta$ describing the off-mass-shell extent of electron
937: in the bound state. For a hydrogenlike atom, we may ascribe
938: $\delta\mu$ to the (minus) binding energy of electron in the Bohr
939: theory:
940: 
941: 
942: \beq
943: \delta\mu=\varepsilon_n=-\frac{Z^2\alpha^2}{2n^2}\mu
944: \label{4-17}
945: \eeq
946: 
947: Then Eq.(\ref{4-16}) gives the value of $\zeta$ for fixed values $Z$
948: and $n$. We will see from the vertex function that these values of
949: $\zeta$ are crucial to the calculation of Lamb shift (sections
950: VII and VIII).
951: 
952: \section{Photon Self-energy}
953: 
954: As discussed in various text books \cite{21,22,23}, we encounter
955: the FDI of vacuum polarization Fig.2(c) as \cite{8:a}:
956: 
957: \beq \Pi_{\mu\nu}(q)=-(-ie)^2Tr\int\frac{d^4\bar{p}}{(2\pi)^4}
958: \gamma_{\mu}\frac{i}{\not\!\!\bar{p}-m}\gamma_{\nu}\frac{i}{\not\!\!\bar{k}+\not\!\!q-m}
959: \label{5-1}
960: \eeq
961: 
962: Here $q$ is the momentum transfer along the photon line and $m$ the mass of electron.
963: Introducing the Feynman parameter $x$ as in previous section and performing a shift in momentum integration:
964: $\bar{p}\rightarrow K=\bar{p}+xq$, we get:
965: 
966: \beq \Pi_{\mu\nu}(q)=-4e^2\int_0^1dx(I_1+I_2)
967: \label{5-2}
968: \eeq
969: 
970: where
971: 
972: \beq
973: I_1=\int\frac{d^4K}{(2\pi)^4}\frac{2K_{\mu}K_{\nu}-g_{\mu\nu}K^2}{(K^2-M^2)^2}
974: \label{5-3}
975: \eeq
976: 
977: with
978: 
979: \beq
980: M^2=m^2+q^2(x^2-x)
981: \label{5-4}
982: \eeq
983: 
984: is quadratically divergent while
985: 
986: \beq
987: I_2=\int\frac{d^4K}{(2\pi)^4}\frac{(x^2-x)(2q_{\mu}q_{\nu}-g_{\mu\nu}q^2)+m^2g_{\mu\nu}}{(K^2-M^2)^2}
988: \label{5-5}
989: \eeq
990: 
991: is only logarithmically divergent like that in Eq.(\ref{4-5}). An
992: elegant way to handle $I_1$, Eq.(\ref{5-3}), is modifying $M^2$ to
993: 
994: \beq
995: M^2(\sigma)=m^2+q^2(x^2-x)+\sigma
996: \label{5-6}
997: \eeq
998: 
999: and differentiating $I_1$ with respect to $\sigma$.
1000: After integration with respect to $K$, we reintegrate it with
1001: respect to $\sigma$ twice, arriving at the limit
1002: $\sigma\rightarrow 0$:
1003: 
1004: \beq
1005: I_1=\frac{ig_{\mu\nu}}{(4\pi)^2}\left\{[m^2+q^2(x^2-x)]\ln\frac{m^2+q^2(x^2-x)}{\mu_3^2}+C_2\right\}
1006: \label{5-7}
1007: \eeq
1008: 
1009: with two arbitrary constant: $C_1=-\ln\mu_3^2$ and $C_2$. Combining
1010: $I_1$ and $I_2$ together, we find:
1011: 
1012: \beq
1013: \Pi_{\mu\nu}(q)=\frac{8ie^2}{(4\pi)^2}(q_{\mu}q_{\nu}-g_{\mu\nu}q^2)\int_0^1dx(x^2-x)
1014: \ln\frac{m^2+q^2(x^2-x)}{\mu_3^2}-\frac{4ie^2}{(4\pi)^2}g_{\mu\nu}C_2
1015: \label{5-8}
1016: \eeq
1017: 
1018: The continuity equation of current induced in the vacuum
1019: polarization \cite{21}
1020: 
1021: \beq
1022: q^{\mu}\Pi_{\mu\nu}(q)=0
1023: \label{5-9}
1024: \eeq
1025: 
1026: is ensured by the factor $(q_{\mu}q_{\nu}-g_{\mu\nu}q^2)$. So we
1027: set $C_2=0$. Consider the scattering between two electrons via the
1028: exchange of a photon with momentum transfer $q\rightarrow0$.
1029: Adding the contribution of $\Pi_{\mu\nu}(q)$ to the tree diagram
1030: amounts to modify the charge square:
1031: 
1032: \beq e^2\longrightarrow e^2_R=Z_3e^2,\ \ \
1033: Z_3=1+\frac{\alpha}{3\pi}(\ln\frac{m^2}{\mu_3^2}-\frac{q^2}{5m^2}+\cdots)
1034: \label{5-10}
1035: \eeq
1036: 
1037: As in Ref.\cite{8:b}, we will set $\mu_3=m$ so that at the Thomson
1038: limit:$\lim_{q\rightarrow0}e_R^2=e^2$. However, for the purpose of
1039: calculating Lamb shift (LS) below, the second term in the
1040: parenthesis of $Z_3$ is important because for a bound state it
1041: contributes a term of effective potential (adding to Coulomb
1042: potential), called the Uehling potential (\cite{23},p.253):
1043: 
1044: \beq
1045: -\frac{4\alpha^2}{15m^2}\delta(\mathbf r)
1046: \label{5-11}
1047: \eeq
1048: 
1049: \section{The Off-Mass-Shell Vertex Function}
1050: 
1051: Consider an electron (see Fig.2(d)) moving in a hydrogen atom,
1052: its momentum changes from $p$ to $p'$ via the scattering by the
1053: proton and an exchange of virtual photon with momentum $k$. The
1054: FDI at one-loop level reads
1055: 
1056: \beq
1057: \Lambda_{\mu}(p',p)=(-ie)^2\int\frac{d^4k}{(2\pi)^4}\frac{-i}{k^2}\gamma_{\nu}
1058: \frac{i}{\not\!\!p'-\not\!\!k'-\mu}\gamma_{\mu}\frac{i}{\not\!\!p-\not\!\!k-\mu}\gamma^{\nu}
1059: \label{6-1}
1060: \eeq
1061: 
1062: However, different from \cite{8:b} and many other literatures, not
1063: only the reduced mass $\mu$ (instead of $m$) of electron is used,
1064: but also a new approach will be adopted. We assume that the
1065: electron is moving off-mass-shell in the sense of (as in section
1066: IV):
1067: 
1068: \beq
1069: p^2=p'^2=\mu^2(1-\zeta)
1070: \label{6-2}
1071: \eeq
1072: 
1073: We still have
1074: 
1075: \beq
1076: p'-p=q,\ \ \ p\cdot q=-\frac{1}{2}q^2
1077: \label{6-3}
1078: \eeq
1079: 
1080: Introducing Feynman parameters $u=x+y$ and $v=x-y$, we perform
1081: a shift in the momentum integration:$k\rightarrow
1082: K=k-(p+q/2)u-(q/2)v$, thus
1083: 
1084: \beq \Lambda_{\mu}=-ie^2[I_3\gamma_{\mu}+I_4]
1085: \label{6-4}
1086: \eeq
1087: 
1088: \bea
1089: I_3&=&\int_0^1du\int_{-u}^udv\int\frac{d^4K}{(2\pi)^4}\frac{K^2}{(K^2-M^2)^3}\label{6-5}\\
1090: M^2&=&[\mu^2(1-\zeta)-\frac{q^2}{4}]u^2+\frac{q^2}{4}v^2+\zeta\mu^2u\label{6-6}\\
1091: I_4&=&\int_0^1du\int_{-u}^udv\int\frac{d^4K}{(2\pi)^4}\frac{A_{\mu}}{(K^2-M^2)^3}\label{6-7}\\
1092: A_{\mu}\!\!\!&=&\!\!\!(4\!\!-\!\!4u\!\!-\!\!2u^2)\mu^2(1\!\!-\!\!\zeta)\gamma_{\mu}\!+\!2i(u^2\!\!-\!\!u)\mu
1093: q^{\nu}\!\sigma_{\mu\nu}
1094: \!-\!(2\!-\!2u\!+\!\frac{u^2}{2}\!-\!\frac{v^2}{2})q^2\gamma_{\mu}\!-\!(2\!+\!2u)v\mu
1095: q_{\mu} \label{6-8} \ena
1096: 
1097: Set $K^2=K^2-M^2+M^2$, then $I_3=I'_3-\frac{i}{32\pi^2}$ and
1098: $I'_3$ is only logarithmically divergent and so can be treated as
1099: in previous sections, yielding:
1100: 
1101: \beq
1102: I'_3=\frac{-i}{(4\pi)^2}\int_0^1du\int_{-u}^udv\ln\frac{M^2}{\mu_1^2}
1103: \label{6-9}
1104: \eeq
1105: 
1106: with $\mu_1$ an arbitrary constant.
1107: 
1108: However, unlike Ref.\cite{8:b} where the calculation was conducted on
1109: the mass-shell, now the off-mass-shell integration in Eq.(\ref{6-9}) can be
1110: performed in the approximation that $\frac{Q^2}{4\mu^2}\ll1$ and
1111: $\zeta\ll1$ ($Q^2=-q^2$, $Q$ is the three-dimensional momentum
1112: transfer) which will be enough to calculate the Lamb shift (LS).
1113: Denoting
1114: 
1115: \beq
1116: a=[\mu^2(1-\zeta)+\frac{Q^2}{4}]u^2+\zeta\mu^2u,\ \ \
1117: b=\frac{Q^2}{4}
1118: \label{6-10}
1119: \eeq
1120: 
1121: we will perform the integration with respect to $v$ and $u$
1122: rigorously:
1123: 
1124: \beq
1125: \int_{-u}^udv\ln(a-bv^2)=2u[\ln\mu^2+\ln
1126: u+\ln[(1-\zeta)u+\zeta]-4u+2\sqrt{\frac{4a}{Q^2}}\ln\frac{1+\sqrt{Q^2/4a}u}{1-\sqrt{Q^2/4a}u}
1127: \label{6-11}
1128: \eeq
1129: 
1130: Expanding the last term and keeping only up to the order of
1131: $\zeta$ and $Q^2/4\mu^2$, we obtain
1132: 
1133: \beq
1134: \int_0^1du\int_{-u}^udv\ln(a-bv^2)\simeq\ln\mu^2-1+\zeta+\frac{Q^2}{6\mu^2}(1-\zeta)
1135: \label{6-12}
1136: \eeq
1137: 
1138: To our great pleasure, throughout the evaluation of $I_4$, there
1139: is no any infrared divergence which would appear in previous
1140: literatures when integrating with respect to $u$ with lower limit
1141: zero. To avoid the infrared divergence, \eg, in \cite{8:b}, a cutoff
1142: was introduced at the lower limit. Now the infrared
1143: divergence disappears due to the existence of off-mass-shell
1144: parameter $\zeta$. For example, we encounter the following
1145: integral, in which no cutoff is needed
1146: ($\lambda=(1-\zeta)+Q^2/4\mu^2\sim1$):
1147: 
1148: \beq
1149: \int_0^1\frac{du}{u+\zeta/\lambda}=\frac{\zeta}{\lambda}-\ln\frac{\zeta}{\lambda}
1150: \label{6-13}
1151: \eeq
1152: 
1153: Hence, after elementary but tedious calculation, we find:
1154: 
1155: \beq\begin{array}{l}
1156: \Lambda_{\mu}(p',p)=\frac{\alpha}{4\pi}[\frac{11}{2}-\ln\frac{\mu^2}{\mu_1^2}-3\zeta+4(1+\zeta)\ln\zeta]\gamma_{\mu}
1157: +\frac{\alpha}{4\pi}\frac{Q^2}{\mu^2}\gamma_{\mu}(\frac{1}{6}+\frac{1}{2}\zeta+\frac{4}{3}\ln\zeta+2\zeta\ln\zeta)\\
1158: +i\frac{\alpha}{4\pi}\frac{q^{\nu}}{\mu}\sigma_{\mu\nu}(1+3\zeta+2\zeta\ln\zeta)
1159: \label{6-14}
1160: \end{array}\eeq
1161: 
1162: \section{Calculation of Lamb Shift as an Off-Mass-Shell Effect at
1163: One-Loop Level}
1164: 
1165: There are three parts in Eq.(\ref{6-14}). The first part in
1166: combination with the vertex $\gamma_{\mu}$ at tree level provides
1167: a renormalization factor as
1168: 
1169: \beq
1170: Z_1^{-1}=1+\frac{\alpha}{4\pi}[\frac{11}{2}-\ln\frac{\mu^2}{\mu_1^2}-3\zeta+4(1+\zeta)\ln\zeta]
1171: \label{7-1}
1172: \eeq
1173: 
1174: Further combination with $Z_2$ in Eq.(\ref{4-10}) and $Z_3$ in
1175: Eq.(\ref{5-10}) leads to a renormalized charge (at one-loop level,
1176: see Fig.2):
1177: 
1178: \beq
1179: e_R=\frac{Z_2}{Z_1}Z_3^{1/2}e
1180: \label{7-2}
1181: \eeq
1182: 
1183: However the Ward identity implies that \cite{21,22,23} \
1184: 
1185: \beq
1186: Z_1=Z_2
1187: \label{7-3}
1188: \eeq
1189: 
1190: Therefore
1191: 
1192: \beq
1193: \alpha_R=\frac{e_R^2}{4\pi}=Z_3\alpha
1194: \label{7-4}
1195: \eeq
1196: 
1197: Note that Ward identity holds not only for an electron on the
1198: mass-shell, but also for off-mass-shell case. Hence for every
1199: bound state in hydrogenlike atom with a definite value of $\zeta$
1200: ($Z_1$ and $Z_2$ are functions of $\zeta$), the arbitrary constant
1201: $\mu_1$ in Eq.(\ref{7-1}) plays a flexible role to guarantee the
1202: validity of Eq.(\ref{7-3}) (other two constants $\mu_2$ and $\mu_3$
1203: had been fixed in Eq.(\ref{4-13}) and (\ref{5-10}) respectively). For further discussion, see section IX.
1204: 
1205: The second part of Eq.(\ref{6-14}) contains $Q^2\gamma_{\mu}$. Just
1206: like the Uehling potential in Eq.(\ref{5-10}) (with $q^2=-Q^2$), it
1207: contributes an effective potential of $\delta$ function type as
1208: 
1209: \beq
1210: \frac{\alpha^2}{\mu^2}[-\frac{1}{6}-\frac{1}{2}\zeta-\frac{4}{3}\ln\zeta-2\zeta\ln\zeta]\delta(\mathbf
1211: r)
1212: \label{7-5}
1213: \eeq
1214: 
1215: Finally, the third part of Eq.(\ref{6-14}) amounts to a modification
1216: of electron magnetic moment in the atom, the gyromagnetic ratio of
1217: electron reads:
1218: 
1219: \beq
1220: g=2[1+\frac{\alpha}{2\pi}(1+3\zeta+2\zeta\ln\zeta)]
1221: \label{7-6}
1222: \eeq
1223: 
1224: We will call the anomalous part of magnetic
1225: moment $a=\frac{\tilde{\alpha}}{2\pi}$, $\tilde{\alpha}=\alpha(1+3\zeta+2\zeta\ln\zeta)$. The radiative correction
1226: on the magnetic moment of an electron has two consequences. One is
1227: a modification to the L-S coupling in a hydrogenlike atom (with
1228: charge number $Z$) \cite{21,22}:
1229: 
1230: \beq
1231: H_{LS}^{rad}=2(\frac{\tilde{\alpha}}{2\pi})\frac{\alpha
1232: Z}{4\mu^2r^3}{\mathbf \sigma\cdot L}
1233: \label{7-7}
1234: \eeq
1235: 
1236: Here the electron mass
1237: has been modified from $m$ (see, \eg, \cite{15}) to $\mu$ which
1238: can be derived from the reduced Dirac equation.
1239: 
1240: Another consequence of anomalous magnetic moment of electron
1241: exhibits itself as an additional potential of $\delta$ function
1242: type like Eq.(\ref{5-11})\cite{21,22}
1243: 
1244: \beq
1245: \frac{Z\alpha\tilde{\alpha}}{2\mu^2}\delta(\mathbf r)
1246: \label{7-8}
1247: \eeq
1248: 
1249: Note that Eqs.(\ref{7-7}) and (\ref{7-8}) are only effective to states
1250: with $L\not=0$ and $S$ state with $L=0$ respectively.
1251: 
1252: Adding the results of Eqs.(\ref{7-7}), (\ref{7-8}) and the sum of Eqs.(\ref{5-11}) and (\ref{7-5})
1253: multiplied by $Z$ together to get all radiative corrections (at one-loop level) on
1254: electron in the hydrogenlike atom, then we get the effective potential as
1255: 
1256: \beq
1257: \begin{array}{ll}
1258: V_{eff}^{rad}&=\frac{Z\alpha^2}{\mu^2}[-\frac{4}{3}\ln\zeta-\frac{1}{2}\zeta-2\zeta\ln\zeta
1259: -\frac{1}{6}-\frac{4}{15}\frac{\mu^2}{m^2}+\frac{1}{2}(1+3\zeta+2\zeta\ln\zeta)
1260: ]\delta(\mathbf r)\\
1261: &+\frac{Z\alpha^2}{4\pi\mu^2r^3}(1+3\zeta+2\zeta\ln\zeta)
1262: {\mathbf \sigma\cdot L}\\
1263: &\simeq\frac{Z\alpha^2}{\mu^2}[-\frac{4}{3}\ln\zeta+\frac{1}{15}+\zeta-\zeta\ln\zeta]\delta(\mathbf
1264: r)+\frac{Z\alpha^2}{4\pi\mu^2r^3}(1+3\zeta+2\zeta\ln\zeta){\mathbf \sigma\cdot L}
1265: \end{array}
1266: \label{7-9}
1267: \eeq
1268: 
1269: where we take $\mu^2/m^2\approx1$ in the Uehling potential
1270: to make the formula simpler for a semi-quantitative calculation.
1271: Eq.(\ref{7-9}) leads to the energy modification of a bound state
1272: (with quantum numbers $n,l,j$) in a hydrogenlike atom:
1273: 
1274: \beq
1275: \delta({\mathbf r})\longrightarrow
1276: |\psi_{ns}(0)|^2=\frac{Z^3\alpha^3}{\pi n^3}\mu^3,\ \ \ (l=0)
1277: \label{7-10}
1278: \eeq
1279: 
1280: \beq
1281: \Delta E^{rad}=\Delta E^{rad}(ns)+\Delta E^{rad}_{LS}
1282: \label{7-11}
1283: \eeq
1284: 
1285: \beq \Delta E^{rad}(ns)=\frac{Z^4\alpha^3}{\pi
1286: n^3}R_y[\frac{8}{3}\ln\frac{1}{\zeta}+\frac{2}{15}+2\zeta(1-\ln\zeta)]\delta_{l0}
1287: \label{7-12}
1288: \eeq
1289: 
1290: \beq
1291: \Delta E^{rad}_{LS}=\frac{Z^4\alpha^3}{\pi
1292: n^3}R_y\frac{1+\zeta(3+2\ln\zeta)}{l(2l+1)(l+1)}\left\{
1293: \begin{array}{ll}
1294: &l,\ \ \ \ \ \ (j=l+1/2)\\
1295: -&(l+1),\ \ \ (j=l-1/2)
1296: \end{array}
1297: \right.
1298: \label{7-13}
1299: \eeq
1300: 
1301: where
1302: 
1303: \beq
1304: R_y=\frac{1}{2}\alpha^2\mu=\frac{\mu}{m}R_\infty
1305: \label{7-14}
1306: \eeq
1307: 
1308: \section{Energy-Level Difference in Hydrogenlike Atom: Theory vs.
1309: Experiment}
1310: 
1311: We will study some energy-level differences near the ground state
1312: of hydrogenlike atoms, where precise experimental data are
1313: available. Theoretically, the energy level is fixed primarily by
1314: the formula derived from the reduced Dirac equation (RDE), \ie,
1315: Eq.(\ref{1-1}) with $m_e$ substituted by $\mu_A$ where the subscript $A$
1316: refers to atom $H$, $D$ or $He^+$, \etal.:
1317: 
1318: \bea
1319: E_A^{RDE}&=&\mu_A[f(n,j)-1]=\frac{m_e}{1+b_A}[f(n,j)-1]\nonumber\\
1320: &=&\frac{1}{1+b_A}(1.2355897\times10^{20})[-\frac{(Z\alpha)^2}{2n^2}-\frac{(Z\alpha)^4}{3n^3}(\frac{1}{j+1/2}-
1321: \frac{3}{4n})-\cdots]\ \ Hz
1322: \label{8-1}
1323: \ena
1324: 
1325: Further recoil corrections Eq.(\ref{1-6}) derived by previous
1326: authors will be divided into two terms:
1327: 
1328: \bea
1329: \Delta E_A^{recoil-1}(n,j)&=&-\frac{\mu^2_A}{2M_A}[f(n,j)-1]^2=-\frac{m_eb_A}{2(1+b_A)^3}[f(n,j)-1]^2
1330: \label{8-2}\\
1331: \Delta E_A^{recoil-2}(n,j,l)&=&\frac{(Z\alpha)^4\mu^3_A}{2n^3m_N^{(A)^2}}
1332: (\frac{1}{j+\frac{1}{2}}-\frac{1}{l+\frac{1}{2}})(1-\delta_{l0})
1333: \label{8-3}
1334: \ena
1335: 
1336: Next comes the radiative correction calculated by QED at one-loop
1337: level, Eq.(\ref{7-11}):
1338: 
1339: \beq \Delta
1340: E_A^{rad}(n,j,l)=\frac{1}{1+b_A}\frac{Z^4}{n^3}(\frac{\alpha^3}{\pi}R_{\infty})
1341: [(-\frac{8}{3}\ln\zeta+\frac{2}{15}+2\zeta(1-\ln\zeta))\delta_{l0}
1342: +\frac{1+\zeta(3+2\ln\zeta)}{2l+1}C_{jl}(1-\delta_{l0})]
1343: \label{8-4}
1344: \eeq
1345: 
1346: where
1347: 
1348: \beq
1349: C_{jl}=\left\{
1350: \begin{array}{ll}
1351: &\frac{1}{l+1},\ \ \ j=l+\frac{1}{2}\\
1352: &-\frac{1}{l},\ \ \  j=l-\frac{1}{2}
1353: \end{array}
1354: \right.
1355: \label{8-5}
1356: \eeq
1357: 
1358: Finally, the finite nucleus size (NS) with radius $r_N^{(A)}$
1359: brings a correction \cite{10}:
1360: 
1361: \beq
1362: \begin{array}{ll}
1363: \Delta E_A^{NS}(n,j)&=\frac{4}{3}(\frac{\mu_A}{m_e})^3\frac{Z^4}{n^3}(\frac{r_N^{(A)}}{a_{\infty}})^2R_{\infty}\delta_{l0}\\
1364: &=(\frac{1}{1+b_A})^3\frac{Z^4}{n^3}(4.386454987\times10^7)[\frac{r_N^{(A)}(fm)}{5.2917725}]^2\delta_{l0}\
1365: \ Hz
1366: \end{array}
1367: \label{8-6}
1368: \eeq
1369: 
1370: As explained in Eq.(\ref{4-16}) with Eq.(\ref{4-17}), the value of
1371: off-mass-shell parameter $\zeta$ in Eq.(\ref{8-4}) can be calculated
1372: from the electron self-energy at one-loop level:
1373: 
1374: \beq
1375: \frac{Z^2\alpha}{n^2}=\frac{1}{2\pi}\frac{(\zeta^{<S>}-2\zeta^{<S>}\ln\zeta^{<S>})}{1+\alpha/3\pi}
1376: \label{8-7}
1377: \eeq
1378: 
1379: where the superscript $<S>$ refers to "self-energy". However, we
1380: may derive the value of $\zeta$ in an alternative way. Divide the
1381: square average of four-dimensional momentum $p$ into two parts:
1382: 
1383: \beq
1384: <p^2>=<E^2>-<{\mathbf p}^2>
1385: \label{8-8}
1386: \eeq
1387: 
1388: where
1389: 
1390: \beq
1391: <E^2>=E^2=(\mu-B)^2\simeq\mu^2-2\mu B,
1392: \label{8-9}
1393: \eeq
1394: 
1395: since the binding energy
1396: 
1397: \beq
1398: B=\frac{Z^2\alpha^2}{2n^2}\mu\ll\mu
1399: \label{8-10}
1400: \eeq
1401: 
1402: The square average of three-dimensional momentum $\mathbf p$,
1403: $<{\mathbf p}^2>$, can be evaluated by the Virial theorem (\eg,
1404: \cite{15}). In a Coulomb field, an electron has potential energy
1405: $V=-\frac{Ze^2}{4\pi r}$ and kinetic energy
1406: $T=\frac{1}{2\mu}{\mathbf p}^2$. Then
1407: 
1408: \bea
1409: <{\mathbf p}^2>&=&2\mu<T>=2\mu[-B-<V>]=2\mu B\nonumber\\
1410: <p^2>&=&\mu^2-4\mu B=\mu^2(1-\frac{4B}{\mu})
1411: \label{8-11}
1412: \ena
1413: 
1414: Comparing Eq.(\ref{8-11}) with $<p^2>=\mu^2(1-\zeta^{<V>})$, we find
1415: 
1416: \beq
1417: \zeta^{<V>}=\frac{4B}{\mu}=\frac{2Z^2\alpha^2}{n^2}
1418: \label{8-12}
1419: \eeq
1420: 
1421: where the superscript $<V>$ refers to "Virial theorem". Table 1
1422: gives the values of $\zeta^{<S>}$ and $\zeta^{<V>}$ with their
1423: logarithm values as well as two kinds of "average", $\zeta^{<S+V>}=\frac{1}{2}(\zeta^{<S>}+\zeta^{<V>})$
1424: and $\zeta^{<SV>}=\sqrt{\zeta^{<S>}\zeta^{<V>}}$, to be used in
1425: Eq.(\ref{8-4}).
1426: 
1427: \vskip 0.5cm
1428: \small
1429: \renewcommand\arraystretch{1.3}
1430: \begin{small}\hspace*{-16mm}\begin{tabular}{|c|c|c|c|c|c|c|c|c|}
1431: \multicolumn{8}{c}{Table 1. Off-mass-shell parameter $\zeta$ and $\ln\zeta$}\\[5pt]
1432:   \hline
1433:   $\frac{Z^2}{n^2}$ & $\zeta^{<S>}\times 10^4$ &-$\ln\zeta^{<S>}$ & $\zeta^{<V>}\times 10^6$ & -$\ln\zeta^{<V>}$ &
1434:   $\zeta^{<S+V>}\times 10^5$ & -$\ln\zeta^{<S+V>}$ & $\zeta^{<SV>}\times 10^5$ & $-\ln\zeta^{<SV>}$ \\
1435:   \hline
1436:   $\frac{1}{16}$ & $1.546093458$ & $8.77461$ & $\frac{\alpha^2}{8}=6.6564192$
1437:    &11.91992886 & $8.0632$ & 9.425609 & 3.2080284 & 10.34727 \\
1438:    \hline
1439:   $\frac{1}{4}$ & $7.446539697$ & 7.20259 & $\frac{\alpha^2}{2}=26.6256771$
1440:   & 10.5336345 & $38.5639$ & 7.860609 & 14.0808 & 8.86816225 \\
1441:    \hline
1442:  1 & $37.73719345$ & 5.57969 & $2\alpha^2=106.502$ &9.147340142 &
1443:  $194.011$ & 6.2450103 & 63.39626 & 7.36351521 \\
1444:   \hline
1445: \end{tabular}\end{small}
1446: \vspace{0.5cm}
1447: \normalsize
1448: 
1449: Now we are in a position to discuss a number of cases:
1450: 
1451: (a) The so-called classic Lamb shift of hydrogen atom was measured
1452: experimentally as \cite{10}:
1453: 
1454: \beq
1455: L_H^{exp}(2S-2P)\equiv E_H(2S_{1/2})-E_H(2P_{1/2})=1057.845\
1456: \ MHz
1457: \label{8-13}
1458: \eeq
1459: 
1460: Theoretically, in this case ($b_H=5.446170255\times10^{-4},\ r_N^H=r_p=0.862 fm$),
1461: Eqs.(\ref{8-1}) and (\ref{8-2}) make no contributions while
1462: Eqs.(\ref{8-3}) and (\ref{8-6}) only contribute
1463: 
1464: \beq
1465: \Delta E_H^{recoil-2}(2S_{1/2}-2P_{1/2})=-E_H^{recoil-2}(2,1/2,1)=-2.16156\ \ kHz
1466: \label{8-14}
1467: \eeq
1468: 
1469: and
1470: 
1471: \beq
1472: \Delta E_H^{NS}(2S-2P)=0.14525347\ \ MHz
1473: \label{8-15}
1474: \eeq
1475: 
1476: respectively. The dominant contribution comes from Eq.(\ref{8-4}).
1477: If using $\zeta^{<S>}$, we obtain
1478: 
1479: \beq\begin{array}{l}
1480: \Delta
1481: E_H^{Rad<S>}(2S-2P)=\frac{1}{1+b_H}\frac{1}{8}(4.06931316\times10^8)
1482: [-\frac{8}{3}\ln\zeta^{<S>}+\frac{7}{15}+3\zeta^{<S>}-\frac{4}{3}\zeta^{<S>}\ln\zeta^{<S>}]\\
1483: =1000.6567\
1484: MHz \label{8-16} \end{array}\eeq
1485: 
1486: If we use another three values of $\ln\zeta$ in Table 1, we get
1487: 
1488: \bea \Delta E_H^{Rad<V>}(2S-2P)&=&1451.7912\ \ MHz \label{8-17}\\
1489: \Delta E_H^{Rad<S+V>}(2S-2P)&=&1089.6513\ \ MHz \label{8-18}\\
1490: \Delta E_H^{Rad<SV>}(2S-2P)&=&1226.0871\ \ MHz
1491:  \label{8-19}
1492: \ena
1493: 
1494: It seems that Eq.(\ref{8-16}) is smaller whereas Eq.(\ref{8-17}) too
1495: large. So as an empirical rule in our semiquantitative
1496: calculation, we may use Eq.(\ref{8-18}) to get
1497: 
1498: \beq L_H^{theor.}(2S-2P)= 1089.651+0.145-0.002=1089.794\ \ MHz
1499: \label{8-20} \eeq
1500: 
1501: which is larger than Eq.(\ref{8-13}) by $3\%$.
1502: 
1503: (b) The Lamb shift of $He^+$ atom has been measured as (quoted
1504: from \cite{27}):
1505: 
1506: \beq L_{He^+}^{exp}(2S-2P)= 14041.13(17)\ \ MHz \label{8-21} \eeq
1507: 
1508: Similar to the case of hydrogen atom but with $Z=2$ and
1509: $b_{He^+}=\frac{m_e}{m_{\alpha}}=0.0001371$, we find
1510: 
1511: \bea \Delta E_{He^+}^{Rad<S>}(2S-2P)&=&1.252680693\times10^{10}\ \ Hz\nonumber\\
1512: \Delta E_{He^+}^{Rad<V>}(2S-2P)&=&2.023083608\times10^{10}\ \ Hz\nonumber\\
1513: \Delta E_{He^+}^{Rad<S+V>}(2S-2P)&=&1.369980830\times10^{10}\ \ Hz\nonumber\\
1514: \Delta E_{He^+}^{Rad<SV>}(2S-2P)&=&1.636521214\times10^{10}\ \ Hz
1515:  \label{8-22}
1516: \ena
1517: 
1518: As in the case of $H$ atom, we take the $<S+V>$ scheme and add
1519: 
1520: \bea \Delta E_{He^+}^{recoil-2}(2S-2P)&=&-2.165\ \ kHz \label{8-23}\\
1521: \Delta E_{He^+}^{NS}(2S-2P)&=&4.514\ \ MHz \label{8-24} \ena
1522: 
1523: ($r_{\alpha}\simeq1.2fm$), to find the theoretical value:
1524: 
1525: \beq L_{He^+}^{theor.}(2S-2P)= 13704.220\ \ MHz \label{8-25} \eeq
1526: 
1527: which is smaller than Eq.(\ref{8-21}) by $2.41\%$.
1528: 
1529: (c) The following energy-level difference is related to the "hyper
1530: Lamb shift (HLS)" \cite{10}:
1531: 
1532: \beq \Delta_H^{exp}\equiv
1533: E_H(4S)-E_H(2S)-\frac{1}{4}[E_H(2S)-E_H(1S)]=4797.338(10)\ \ MHz
1534: \label{8-26} \eeq
1535: 
1536: Theoretically, now Eq.(\ref{8-1}) makes the main contribution:
1537: 
1538: \beq \Delta
1539: E_H^{RDE}[(4S)-\frac{5}{4}(2S)+\frac{1}{4}(1S)]=3923.95\ \ MHz
1540: \label{8-27} \eeq
1541: 
1542: (The notation in parenthesis is self-evident). Eq.(\ref{8-2}) and Eq.(\ref{8-4})
1543: contribute
1544: 
1545: \beq \Delta E_H^{recoil-1}[(4S)-\frac{5}{4}(2S)+\frac{1}{4}(1S)]=
1546: -4.186\ \ MHz \label{8-28} \eeq
1547: 
1548: and
1549: 
1550: \bea \Delta E_H^{Rad<S>}&=&451.229097\ \ MHz\nonumber\\
1551: \Delta E_H^{Rad<S+V>}&=&529.288296\ \ MHz\nonumber\\
1552: \Delta E_H^{Rad<SV>}&=&675.907131\ \ MHz\nonumber\\
1553: \Delta E_H^{Rad<V>}&=&903.266275\ \ MHz
1554:  \label{8-29}
1555: \ena
1556: 
1557: respectively. Adding a small contribution from Eq.(\ref{8-6})
1558: 
1559: \beq \Delta E_H^{NS}[(4S)-\frac{5}{4}(2S)+\frac{1}{4}(1S)]=
1560: 0.1270967854\ \ MHz \label{8-30} \eeq
1561: 
1562: we get
1563: \bea
1564: \Delta_H^{Theor.<S>}&=&4371.120197\ \ MHz\nonumber\\
1565: \Delta_H^{Theor.<S+V>}&=&4449.179396\ \ MHz\nonumber\\
1566: \Delta_H^{Theor.<SV>}&=&4595.798231\ \ MHz\nonumber\\
1567: \Delta_H^{Theore.<V>}&=&3923.95-4.186+903.266275+0.1271=4823.1574\ \ MHz \label{8-31}
1568: \ena
1569: 
1570: The $<V>$ scheme is only larger than Eq.(\ref{8-26}) by $0.54\%$.
1571: All other schemes would be too small. So we guess that
1572: for $S$ states $<V>$ scheme is better than $<S>$ scheme.
1573: 
1574: (d) The following energy-level difference was also measured as
1575: \cite{10}:
1576: 
1577: \beq {\Delta'}_H^{exp}\equiv
1578: E_H(4D_{5/2})-E_H(2S)-\frac{1}{4}[E_H(2S)-E_H(1S)]=6490.144(24)\ \
1579: MHz \label{8-32} \eeq
1580: 
1581: Theoretically, Eq.(\ref{8-1}) also makes the main contribution:
1582: 
1583: \beq \Delta
1584: E_H^{RDE}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]=5747.92\ \
1585: MHz \label{8-33} \eeq
1586: 
1587: while
1588: 
1589: \beq \Delta
1590: E_H^{recoil-1}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]=-4.18611\
1591: \ MHz \label{8-34} \eeq
1592: 
1593: \beq \Delta
1594: E_H^{recoil-2}[(4D_{5/2})]=\alpha^4m_e(5.446170255\times10^{-4})^2(\frac{1}{3}-\frac{2}{5})=-6.9283\
1595: \ kHz \label{8-35} \eeq
1596: 
1597: are all small, we will have
1598: 
1599: \bea {\Delta'}
1600: E_H^{rad<S>}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]&=&
1601: 302.088631\ \ MHz\nonumber\\
1602: {\Delta'}E_H^{rad<V>}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]&=&
1603: 700.843464\ \ MHz\nonumber\\
1604: {\Delta'}E_H^{rad<S+V>}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]&=&
1605: 369.124660\ \ MHz\nonumber\\
1606: {\Delta'}E_H^{rad<SV>}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]&=&
1607: 500.131264\ \ MHz \label{8-36} \ena
1608: 
1609: Finally, the nucleus size effect gives
1610: 
1611: \beq
1612: {\Delta'}E_H^{NS}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]=
1613: 11.62027752\times10^5(\frac{1}{4}-\frac{5}{4}\times\frac{1}{8})=0.10894\
1614: \ MHz \label{8-37} \eeq
1615: 
1616: In sum, we have
1617: 
1618: \bea
1619: {\Delta'}_H^{<S>}&=&{\Delta'}E_H^{RDE}+{\Delta'}E_H^{recoil-1}+{\Delta'}E_H^{recoil-2}+{\Delta'}
1620: E_H^{rad<S>}+{\Delta'} E_H^{rad<NS>}=6045.925\ \ MHz\nonumber\\
1621: {\Delta'}_H^{<V>}&=&6444.679\ \ MHz\nonumber\\
1622: {\Delta'}_H^{<S+V>}&=&6112.961\ \ MHz\nonumber\\
1623: {\Delta'}_H^{<SV>}&=&6243.967\ \ MHz \label{8-38}
1624: \ena
1625: 
1626: which are smaller than the experimental value (\ref{8-32}) by
1627: $6.8\%,\  0.7\%,\  5.8\%$ and $3.8\%$ respectively.
1628: .
1629: 
1630: (e) Experimentally, the combination of Eq.(\ref{8-26}) with
1631: Eq.(\ref{8-32}) yields:
1632: 
1633: \beq {\Delta''}_H^{exp}\equiv E(4D_{5/2})-E(4S_{1/2})=1692.806\ \
1634: MHz \label{8-39} \eeq
1635: 
1636: Then, theoretically, we have
1637: 
1638: \bea
1639: {\Delta''}_H^{RDE}(4D_{5/2}-4S)&=&1.823886903\times10^9\ \ Hz \label{8-40}\\
1640: {\Delta''}_H^{recoil-1}(4D_{5/2}-4S)&=&1.1008\ \ Hz \label{8-41}\\
1641: {\Delta''}_H^{recoil-2}(4D_{5/2}-4S)&=&-6.9283\ \ kHz \label{8-42}\\
1642: {\Delta''}_H^{NS}(4D_{5/2}-4S)&=&-0.0181605862\ \ MHz \label{8-43}
1643: \ena
1644: 
1645: and
1646: 
1647: \bea
1648: {\Delta''}_H^{rad<S>}(4D_{5/2}-4S)&=&-149.1404661\ \ MHz \label{8-44}\\
1649: {\Delta''}_H^{rad<V>}(4D_{5/2}-4S)&=&-202.4228107\ \ MHz \label{8-45}\\
1650: {\Delta''}_H^{rad<S+V>}(4D_{5/2}-4S)&=&-160.1636366\ \ MHz \label{8-46}\\
1651: {\Delta''}_H^{rad<SV>}(4D_{5/2}-4S)&=&-175.7758676\ \ MHz \label{8-47}
1652: \ena
1653: 
1654: Altogether, we have
1655: 
1656: \bea
1657: {\Delta''}_H^{theore.<S>}(4D_{5/2}-4S)&=&1674.721349\ \ MHz\nonumber\\
1658: {\Delta''}_H^{theore.<S+V>}(4D_{5/2}-4S)&=&1663.716339\ \ MHz\nonumber\\
1659: {\Delta''}_H^{theore.<SV>}(4D_{5/2}-4S)&=&1648.104108\ \ MHz\nonumber\\
1660: {\Delta''}_H^{theore.<V>}(4D_{5/2}-4S)&=&1621.439105\ \ MHz
1661: \label{8-48}
1662: \ena
1663: 
1664: which are smaller than Eq.(\ref{8-39}) by $1.1\%,\ 1.7\%,\ 2.6\%$ and $4.2\%$ respectively.
1665: 
1666: (f) It's time to go back to the precision data of $2S-1S$
1667: transition in hydrogen atom as discussed in section II. Rewrite
1668: Eq.(\ref{2-1}) as (see also \cite{43}):
1669: 
1670: \beq \Delta E_H^{exp}(2S-1S)=2.46606141318734\times10^{15}\ \ Hz
1671: \label{8-49} \eeq
1672: 
1673: Theoretically, we have [see Eq.(\ref{2-4})]:
1674: 
1675: \beq \Delta E_H^{RDE}(2S-1S)=2.466067984\times10^{15}\ \ Hz
1676: \label{8-50} \eeq
1677: 
1678: \beq \Delta E_H^{recoil-1}(2S-1S)=22.32598676\ \ MHz
1679: \label{8-51}
1680: \eeq
1681: 
1682: \bea \Delta E_H^{rad<S>}(2S-1S)&=&-5142.081146\ \ MHz\nonumber\\
1683: \Delta E_H^{rad<S+V>}(2S-1S)&=&-5765.958928\ \ MHz\nonumber\\
1684: \Delta E_H^{rad<SV>}(2S-1S)&=&-6835.535314\ \ MHz\nonumber\\
1685: \Delta E_H^{rad<V>}(2S-1S)&=&-8541.095068\ \ MHz
1686:  \label{8-52}
1687: \ena
1688: 
1689: \beq
1690: \Delta E_H^{NS}(2S-1S)=11.62027752\times10^5(\frac{1}{8}-1)=-1.016774283\
1691: \ MHz
1692:  \label{8-53}
1693: \eeq
1694: 
1695: If taking the value of $\Delta E_H^{rad}(2S-1S)$, we get
1696: 
1697: \bea
1698: \Delta E_H^{theore.<S>}(2S-1S)&=& 2.466062836\times10^{15} \ \ Hz\nonumber\\
1699: \Delta E_H^{theore.<S+V>}(2S-1S)&=& 2.466062239\times10^{15} \ \ Hz\nonumber\\
1700: \Delta E_H^{theore.<SV>}(2S-1S)&=& 2.466061169\times10^{15} \ \ Hz\nonumber\\
1701: \Delta E_H^{theore.<V>}(2S-1S)&=& 2.466059464\times10^{15}\ \ Hz \label{8-54}
1702: \ena
1703: 
1704: They are larger than Eq.(\ref{8-49}) by $1450\ MHz,\ 826\ MHz$ and smaller than Eq.(\ref{8-49})
1705: by $244\ MHz,\ 1949\ MHz$ respectively. Or, their discrepancies are
1706: $+5.9\times10^{-7},\ +3.3\times10^{-7},\ -1.0\times10^{-7},\ -7.9\times10^{-7}$, respectively.
1707: This discrepancy is basically stemming from
1708: the uncertainty in the calculation of $\Delta E_H^{rad}(2S-1S)$.
1709: 
1710: (g) Let us turn to the isotope-shift of $2S-1S$ transition.
1711: Rewrite Eq.(\ref{2-2}) as
1712: 
1713: \beq \Delta E_{D-H}^{exp}(2S-1S)=6.70994337\times10^{11}\ \ Hz
1714: \label{8-55} \eeq
1715: 
1716: Theoretically, rewrite Eqs.(\ref{2-6}) and (\ref{2-7}) as
1717: 
1718: \beq \Delta E_{D-H}^{RDE}(2S-1S)=6.7101527879\times10^{11}\ \ Hz
1719: \label{8-56} \eeq
1720: 
1721: and
1722: 
1723: \beq \Delta E_{D-H}^{recoil-1}(2S-1S)=-11.176\ \ MHz \label{8-57}
1724: \eeq
1725: 
1726: \bea \Delta E_{D-H}^{rad<S>}(2S-1S)&=&-1.399158\ \ MHz\nonumber\\
1727: \Delta E_{D-H}^{rad<V>}(2S-1S)&=&-2.324028\ \ MHz\nonumber\\
1728: \Delta E_{D-H}^{rad<S+V>}(2S-1S)&=&-1.568915\ \ MHz\nonumber\\
1729: \Delta E_{D-H}^{rad<SV>}(2S-1S)&=&-1.859945\ \ MHz
1730:  \label{8-58}
1731: \ena
1732: 
1733: \beq \Delta E_{D-H}^{NS}(2S-1S)=-5.11384949\ \ MHz \label{8-59}
1734: \eeq
1735: 
1736: Altogether, we find [using $<V>$ scheme in Eq.(\ref{8-58})]:
1737: 
1738: \beq \Delta
1739: E_{D-H}^{theore.<V>}(2S-1S)=6.709966701\times10^{11}\ \ Hz
1740: \label{8-60} \eeq
1741: 
1742: which is larger than Eq.(\ref{8-55}) by $2.333\ MHz$ or only
1743: $3.5\times10^{-6}$. Evidently, even Eq.(\ref{8-56}) solely
1744: deviates from Eq.(\ref{8-55}) by $3\times10^{-5}$ only. And as
1745: expected, the different schemes for $\Delta E_{D-H}^{rad}(2S-1S)$
1746: have little influence on the theoretical value, because any one of
1747: Eq.(\ref{8-58}) is much smaller than the nucleus size effect
1748: Eq.(\ref{8-59}) ($r_N^D=r_d=2.115 fm$).
1749: 
1750: (h) Finally, the so-called absolute Lamb-shift of $1S$ state in
1751: hydrogen atom was determined by Weitz \etal \cite{10}
1752: from the measured value Eq.(\ref{8-26}) or (\ref{8-32}). In our
1753: notation, using Eq.(\ref{8-32}), we will write it as follows:
1754: 
1755: \beq
1756: \begin{array}{ll}
1757: L_H(1S)=&4\{{\Delta'}_H^{exp}-\Delta
1758: E_H^{RDE}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]-\Delta
1759: E_H^{recoil-1}[(4D_{5/2})-\frac{5}{4}(2S)+\frac{1}{4}(1S)]\\
1760: &-\Delta
1761: E_H^{recoil-2}(4D_{5/2})+\frac{5}{4}L_H(2S)-L_H(4D_{5/2})\}
1762: \end{array}
1763: \label{8-61} \eeq
1764: 
1765: Here the Lamb shift of $2S$ state $L_H(2S)$ can be determined from
1766: the experimental value of Eq.(\ref{8-13}) with $L_H(2P_{1/2})$
1767: being calculated from Eq.(\ref{8-4}):
1768: 
1769: \beq L_H(2S)=L_H^{exp}(2S-2P)-\Delta
1770: E_H^{recoil-2}(2S-2P_{1/2})+L_H(2P_{1/2})=1040.901\ \ MHz
1771: \label{8-62} \eeq
1772: 
1773: And $L_H(4D_{5/2})$ can also be calculated from Eq.(\ref{8-4}), so
1774: 
1775: \beq L_H(1S)=8188.478\ \ MHz \label{8-63} \eeq
1776: 
1777: which is in agreement with $8172.874(60)\ MHz$ given by \cite{10}
1778: within an accuracy $\lesssim 0.2\%$. If we use Eq.(\ref{8-26}) to
1779: derive $L_H(1S)$, we would have to calculate $L_H(4S)$ which is
1780: much larger than $L_H(4D_{5/2})$ and its derivation from
1781: Eq.(\ref{8-4}) seems not reliable. Similarly, the theoretical
1782: value of $L_H(1S)$ turns out to be
1783: 
1784: \beq L_H^{theore.}(1S)=\Delta E_H^{rad}(1S)+\Delta E_H^{NS}(1S)
1785: \label{8-64} \eeq
1786: 
1787: with $\Delta E_H^{NS}(1S)=0.14525347\ MHz$. However, the value of $\Delta
1788: E_H^{rad}(1S)$ strongly depends on the scheme we used in
1789: Eq.(\ref{8-4}), which must
1790: be narrowed in a high-loop calculation.
1791: The theoretical prediction was given in \cite{27} as:
1792: \begin{equation}\label{8-65}
1793:     L_H^{theor.}(1S)=8172754(14)(32)\ kHz
1794: \end{equation}
1795: Further discussions can be found in Refs. \cite{5,44,45}.
1796: 
1797: \section{Summary and Discussion}
1798: 
1799: The remarkable progress of the experimental research on
1800: energy-level differences in hydrogenlike atoms has been making
1801: this field an ideal theoretical laboratory for physics:
1802: 
1803: (a) The inevitable and successful use of reduced Dirac equation
1804: (RDE) to hydrogenlike atoms, especially to the isotope-shift of
1805: $2S-1S$ transition as reflected by Eqs.(\ref{8-55}) through
1806: (\ref{8-60}), is by no means an accidental fortune. It implies
1807: that the argument in section III for introducing RDE, Eq.(\ref{3-21}),
1808: is correct to a high accuracy. In particular, the basic principle of
1809: invariance under space-time inversion Eq.(\ref{3-22}) (with original mass $m$) could remain
1810: valid even for a noninertial frame. This implication has a far-reaching consequence that a
1811: generalization at the above symmetry to a localized curved
1812: space-time may be served as a possible road to quantize the
1813: general theory of relativity \cite{16}.
1814: 
1815: However, there are two realizations of potential $V$ under the
1816: space-time inversion, Eq.(\ref{3-15})("scalar" type) and
1817: Eq.(\ref{3-22})("vector" type). While Eq.(\ref{3-22}) does
1818: dominant in an atom like $H$ with $m_p\gg m_e$, the remaining
1819: discrepancy of $2.333\ MHz$ between theory and experiment
1820: [Eq.(\ref{8-60}) versus Eq.(\ref{8-55})] strongly hints that an
1821: important and subtle effect had been ignored. (To consider the contribution of
1822: the deuteron polarizability merely accounts for about $20\ kHz$
1823: \cite{6}). We think what neglected must be a tiny excitation of antiparticle
1824: field in the nucleus due to its interaction with electron in the CMCS. So when we reduce the degrees of freedom
1825: of two-body system from two to
1826: one, the RDE should be modified
1827: to take account of the tiny mixture of "scalar" potentials (see the page note after Eq.(\ref{3-22})).
1828: We don't know how to improve $RDE$ yet.
1829: However, an experimental evidence for the above conjecture could be the
1830: following prediction: The discrepancy between present theory (with RDE) and
1831: experiment must be smaller for the isotope shift in $2S-1S$
1832: transition of atoms $^4He$ and $^3He$ than that of atoms $H$ and $D$.
1833: 
1834: Recently, by using Dirac's method, Marsch rigorously solved the
1835: hydrogen atom as a two-Dirac particle system bound by Coulomb
1836: force \cite{34}. His solutions are composed of positive and
1837: negative pairs, corresponding respectively to hydrogen and
1838: anti-hydrogen as expected. However, surprisingly, in the hydrogen
1839: spectrum, besides the normal type-1 solution with reduced mass $\mu$, there is another
1840: anomalous type$-2$ solution with energy levels:  ${E'}_n=
1841: Mc^2-2\mu c^2+\frac{1}{2}\mu c^2(\frac{\alpha}{n})^2+\cdots\
1842: (n=1,2,\ldots)$ and "strange enough, the type$-2$ ground state
1843: $(n=1)$ does not have lowest energy but the continuum
1844: $(n=\infty)$". In our opinion, based on what we learnt from the
1845: Dirac equation and RDE, these anomalous solutions imply a positron
1846: moving in the field of proton. So all discrete states with energy
1847: ${E'}_n$ are actually unbound, they should be and can be ruled out
1848: in physics either by the "square integrable condition" or the
1849: "orthogonality criterion" acting on their rigorous wave functions
1850: (for one-body Dirac equation, see \cite{35}, also p.$28-31$, $50$ of \cite{36}). On the other
1851: hand, all continuum states ($n=\infty$) with energies lower than
1852: $Mc^2-2\mu c^2$ correspond to scattering wave functions with
1853: negative phase shifts , showing the repulsive force between
1854: positron and proton. (see \cite{37}, section 1.5 in \cite{36} or
1855: section 9.5 of \cite{15}). Marsch's discovery precisely reflects
1856: two things: (a) the negative energy state of a particle just
1857: describes its antiparticle state. (b) The Coulomb potential allows
1858: a complete set of solutions comprising of two symmetric
1859: sectors,hydrogen and antihydrogen.In the hydrogen sector, the
1860: proton remains unchanged regardless of the changing process of
1861: electron into positron under the Coulomb interaction.
1862: 
1863: The above particle-antiparticle symmetry (including Eq.(\ref{3-22}) showing the unequal treatment between electron and
1864: nucleus), together with the parity symmetry, is hidden in the Dirac's
1865: four-component theory in covariant form so they were overlooked to some extent in the past. The advantage
1866: or flexibility of two-component noncovariant form of Dirac equation or RDE (as discussed in this
1867: paper) lies in the fact that the above two symmetries become accurate and so easily to be extended
1868: (or violated) in an explicit manner. For completeness, let us stress again that for antiparticle, one should
1869: use the momentum and energy operators being ${\mathbf p}_c=i\nabla$ and $E_c= -i\frac{\partial}{\partial t}$
1870: versus ${\mathbf p}=-i\nabla$ and $E=i\frac{\partial}{\partial t}$ for particle as required by the
1871: space-time inversion symmetry. The historical mission of the conception to imagine the positron as a
1872: "hole" in the sea of negative energy electrons is already over. Since the CPT invariance had been further verified \cite{39},
1873: the relation between a particle $|a\rangle$ and its antiparticle $|\bar{a}\rangle$ is well-established as: \footnotemark[4]
1874: \footnotetext[4]{To our knowledge, the correct definition, Eq.(\ref{167}), was first given by T. D. Lee and C. S. Wu at
1875: Ann. Rev. Nucl. Sci. {\bf 15}, 381(1965). See also G. J. Ni at J. Fudan Univ. (Natural Science) No.3-4, 125(1974).}
1876: 
1877: \begin{equation}\label{167}
1878: |\bar{a}\rangle=CPT|a\rangle
1879: \end{equation}
1880: with their wave-functions (in free motion) being respectively:
1881: \begin{equation}\label{awave}
1882: \langle{\bf x},t|a\rangle\sim \exp[\frac{i}{\hbar}({\bf p}\cdot{\bf x}-Et)]
1883: \end{equation}
1884: \begin{equation}\label{abar}
1885: \langle{\bf x},t|\bar{a}\rangle\sim \exp[-\frac{i}{\hbar}({\bf p}\cdot{\bf x}-Et)]
1886: \end{equation}
1887: Note that in Eqs.(\ref{awave}) and (\ref{abar}), they have the same momentum $\bf p$ and positive energy $E$.
1888: Either a newly defined space-time inversion (${\bf x}\to -{\bf x},\,t\to -t$)
1889: or a simple change of $i\to -i$ will transform Eq.(\ref{awave}) into Eq.(\ref{abar}) (or vice versa).
1890: 
1891: 
1892: (b) Throughout this paper, the electron bound in an atom
1893: is just treated like a stationary "ball" with nucleus at its center and
1894: having a (Bohr) radius ($\sim 1/\alpha m_e$). However, it is in an
1895: off-mass-shell state (In some sense, our atom model is just the opposite to J. J. Thomson's atom model
1896: 100 years ago). In fact, the electron's mass is reduced
1897: suddenly from $m_e$ to $\mu$ in the RMCS when it is captured by a
1898: nucleus at the far remote orbit with quantum number
1899: $n\longrightarrow\infty$ and further reduced to
1900: $\mu+\delta\mu\simeq\mu-\frac{Z^2\alpha^2}{2n^2}\mu$ until $n$
1901: decreasing to the lowest limit $n=1$. The Lamb shift should be
1902: viewed as a further modification on the mass of an off-mass-shell
1903: electron due to radiative correction.
1904: 
1905:  Notice that the parameter $Q^2$ in the vertex function,
1906:  Eq.(\ref{6-14}), means the square of (three-dimensional) momentum
1907:  transfer when a free electron is on its mass-shell and collides
1908:  with some other particle as discussed in Ref.\cite{8:b}. By contrast, now
1909:  $Q^2$ exhibits itself as an effective potential of
1910:  $\delta$-function type exerted by the nucleus to the bound (and
1911:  so off-mass-shell) electron as shown by Eq.(\ref{7-5}). To bind an electron to a nucleus is a
1912:  nonperturbative effect. Hence we can understand why the discrepancy between $\zeta^{<S>}$
1913:  (calculated by perturbative QED at one-loop order) and $\zeta^{<V>}$ (evaluated via nonperturbative
1914:  Virial theorem) is so large. Fortunately, they lead to discrepancies in the calculated values of Lamb shift being not so large as
1915:  shown in Section VIII. When $\zeta^{<V>}$ or $\zeta^{<S+V>}$ (or $\zeta^{<SV>}$) is substituted into the Eq.(\ref{8-4})
1916:  which is derived from perturbative ($L=1$) theory, we should always be aware of some theoretical inconsistency in such a
1917:  semi-empirical treatment. But as a whole, we believe that the concept of Lamb shift as an off-mass-shell effect in covariant QED
1918:  is basically correct.
1919: 
1920:  (c) For a free on-mass-shell electron, its charge square $e_R^2$
1921:  will increase with the increase of $Q^2$ as shown by
1922:  Eq.(\ref{5-10}) (with $\mu_3=m_e,\ q^2=-Q^2$) and was calculated
1923:  in detail in \cite{8:b}, coinciding with the experimental data.
1924:  Note that, however, the Ward identity $Z_1=Z_2$ had been
1925:  used. An interesting question arises for a bound electron: as its
1926:  $e_R^2$ is not a function of $Q^2$, will $e_R^2$ change with the
1927:  variation of the quantum number $n$? To answer this question, let us
1928:  put Ward identity aside for a while and write down the
1929:  renormalized $\alpha_R=\frac{e_R^2}{4\pi}$ as
1930: 
1931:  \beq
1932: \alpha_R=\frac{Z_2^2}{Z_1^2}Z_3\alpha\longrightarrow\frac{Z_2^2}{Z_1^2}\alpha
1933: \label{9-1} \eeq
1934: 
1935: Let us work in the CMCS, so $Z_2=\frac{1}{1-B}$ and $B$ is shown
1936: in Eq.(\ref{4-8}) but with $\mu$ replaced by $m_e=m$. Similarly,
1937: $Z_1$ is given by the first part of Eq.(\ref{6-14}) with
1938: $\mu\longrightarrow m$:
1939: 
1940: \beq Z_1\simeq
1941: 1+\frac{\alpha}{4\pi}[\frac{11}{2}-\ln\frac{m^2}{\mu_1^2}-3\xi+4(1+\xi)\ln\xi]
1942: \label{9-2} \eeq
1943: 
1944: where the off-mass-shell parameter $\xi$ in CMCS is defined by
1945: 
1946: \beq
1947: p^2=m^2(1-\xi)=m^2(1-\eta-\zeta')=m^2(1-\eta)-m^2\zeta'=\mu^2-m^2\zeta'=\mu^2(1-\zeta)
1948: \label{9-3} \eeq
1949: 
1950: with
1951: 
1952: \beq m^2(1-\eta)=\mu^2, \ \ \eta=1-\frac{\mu^2}{m^2},\ \
1953: \zeta'=\frac{\mu^2}{m^2}\zeta \label{9-4} \eeq
1954: 
1955: and $\zeta$ is exactly that in Eq.(\ref{7-1}). If we ignore the
1956: dependence of $(1-B)$ on $\zeta$, Eq.(\ref{9-1}) would give
1957: ($\zeta\ll1$):
1958: 
1959: \beq
1960: \alpha_R=\alpha[1+\frac{2\alpha}{\pi}\ln(1+\frac{\zeta}{\eta})]
1961: \label{9-5} \eeq
1962: 
1963: after renormalization by adjusting the arbitrary constant $\mu_1$
1964: so that
1965: 
1966: \beq \alpha_R|_{\zeta\rightarrow0}=\alpha \label{9-6} \eeq
1967: 
1968: which connects to the Thomson limit
1969: $\alpha_R|_{Q\rightarrow0}=\alpha$ for a free electron
1970: continuously but not smoothly. Then for two lowest bound
1971: states with $n=1$ and $n=2$, we would have (in $<V>$ scheme):
1972: 
1973: \beq \alpha_R|_{n=1}\simeq\alpha(1.000433832),\ \
1974: \alpha_R|_{n=2}\simeq\alpha(1.0001123)\label{9-7} \eeq
1975: 
1976: This would modify the Bohr energy level in hydrogenlike atom A to
1977: 
1978: \beq \tilde{E}_A^{Bohr}(n)=-\frac{Z^2\alpha_n^2}{2n^2}\mu_A
1979: \label{9-8} \eeq
1980: 
1981: and make an extra contribution to the isotope-shift as
1982: 
1983: \beq \Delta \tilde{E}_{D-H}^{Bohr}(2S-1S)\simeq726\ \ MHz
1984: \label{9-9} \eeq
1985: 
1986: which is definitely excluded by the experiment. Hence the above
1987: consideration from Eq.(\ref{9-1}) till Eq.(\ref{9-9}) is wrong. We
1988: learn concretely once again that the Ward identity $Z_1=Z_2$ is valid not
1989: only for an electron on its mass-shell, but also for
1990: off-mass-shell case. Thus we use the same value of $\alpha$
1991: throughout the whole calculation.
1992: 
1993: (d) In Ref.\cite{8:b}, using our RRM and new renormalization group equation (RGE) for QCD derived from it and keeping all masses of 6 quarks ($m_c=1.031\; GeV$, $m_b=4.326\; GeV$,
1994: $m_t=175\; GeV$, $m_s=200 \;MeV$, $m_u=8\; MeV$, $m_d=10\; MeV$), we calculated the strong coupling constant $\alpha_{s_i}(Q)$ for $i=u,d,s,c,b$ respectively.
1995: Their running curves (starting from the common renormalization point $\alpha_s(M_Z)=0.118$) follow the trend of experimental data
1996: (as shown on p.158 of \cite{39}) quite well but separate at the low $Q$ region. Interesting enough, each of them rises to a maximum
1997: $\alpha_{s_i}^{max}$ and then suddenly drops to zero at $Q=0$ corresponding to a threshold energy scale $E_i^{th}$ which could
1998: be explained as the excitation energy scale for breaking the quark pair. For example, we find $E_b^{th}=1.13\; GeV$ which is
1999: just the hadronization energy scale of Upsilon $\Upsilon(b\bar{b})$ against its dissociation into two bosons. Experimentally,
2000: $M(\Upsilon(4s))-M(\Upsilon)=1.12\;GeV$ and $\Upsilon(4s)\to B^+B^-$ or $B^0\bar{B}^0$. similarly, $E_c^{th}= 0.398\;GeV$
2001: while $M(\psi(3770))-M(\psi(3097))=673 \;MeV$ and $\psi(3770)\to D^+D^-$ or $D^0\bar{D}^0$. it seems that
2002: $E_s^{th}\sim 90\,MeV$ and $E_{u,d}^{th}\sim 0.4\,MeV$ are not so reliable but still reasonable.
2003: 
2004: Actually, our calculation on QCD is backed by that on QED. In [8]b, using our RRM and improved RGE, keeping all contributions from 9
2005: charged leptons and quarks we were able to calculate the running fine-structure constant $\alpha_R(Q)$ from the renormalization point
2006: $\alpha_R(Q)|_{Q=0}=\alpha=(137.036)^{-1}$ until it coincides with the experimental value of $\alpha_{exp}(M_Z)=(128.89)^{-1}$.
2007: We fitted quark's masses as mentioned above and found no further room left for extra charged elementary particles (say, of 4th
2008: generation).
2009: 
2010: (e) In 1989, we had estimated the upper and lower bounds on Higgs mass $M_H$ by using a nonperturbative approach in QFT --- the Gaussian effective
2011: potential (GEP) method, yielding \cite{40}:
2012: \begin{equation}\label{}
2013: 76\; GeV <M_H<170\; GeV
2014: \end{equation}
2015: Like many other authors, we were bothered a lot by divergences. After a deeper study on the $\lambda\phi^4$ model by using our new RRM
2016: \cite{8:a}, we restudied this problem by combination GEP with RRM, yielding\cite{24}:
2017: \begin{equation}\label{higgs}
2018: M_H=138\; GeV
2019: \end{equation}
2020: This is not a upper or lower bound but a prediction based on the input of experimental data:
2021: \begin{equation}\label{}\begin{array}{l}
2022: M_W=80.359 \;GeV,\;M_Z=91.1884 \;GeV, \alpha^{-1}=\dfrac{4\pi}{g^2\sin^2\theta_W}=128.89,\;\\
2023: \sin^2\theta_W=\dfrac{{g'}^2}{g^2+{g'}^2}=0.2317
2024:  \end{array}\end{equation}
2025: where $\theta_W$ is the weak mixing (Weinberg) angle. Because of getting rid of all divergences, our calculation is well under
2026: control at every step. As now the search for Higgs particle becomes so urgent experimentally but the theoretical estimation about
2027: its mass still remains uncertain\cite{39}, we think our approach with the prediction (\ref{higgs}) deserves to be reconsidered.
2028: 
2029: (f) Moreover, our RRM can be used in $D+1$ space-time without
2030: limitation on the space dimension $D$. A detailed analysis of
2031: sinh(sine)-Gordon models with $D=1,2$ and $3$ (also using GEPM) is
2032: given by Ref.\cite{25}. Another example is again the Lamb shift but
2033: calculated by QED in noncovariant form and by using RRM similar to
2034: that in this paper, see Appendix (\cite{26}, see also the Appendix 9A in
2035: \cite{15}). The theoretical value (A.20) seems better than (\ref{8-20}), showing that for
2036: dealing with the Lamb shift, the noncovariant theory may be more suitable than the
2037: covariant one at least in the lowest order.
2038: 
2039: (g) Previously, the theories for Lamb shift or generally for
2040: calculating energy levels in hydrogenlike atoms are rather complicated
2041: as reviewed in refs \cite{1,2} and \cite{27}, some of them have been discussed
2042: in the Appendix of this paper. For further clarity, let us try to
2043: summarize the main obstacles, or challenges in four points:\\
2044: (1) Different masses of nuclei must be taken into account;\\
2045: (2) Relativistic effects of the electron (not nucleus) are important;\\
2046: (3) In calculating radiative corrections, the divergence becomes
2047: severer and severer with the increase of loop number;\\
2048: (4) Since nuclei's properties are different from one atom to another,
2049: to treat each atom as a two-body system individually would be a
2050: daunting task, it couldn't be rigorous eventually too. This can be
2051: clearly seen from the recent work by Marsch \cite{34}.
2052: 
2053:    Facing these challenges and learning from lessons and experiences
2054: of previous authors, we see that the clue point is to replace the
2055: electron mass $m$ by reduced mass $\mu$ and work in the noninertial frame
2056: (RMCS) throughout the entire calculation. As is well known, this can
2057: be handled in nonrelativistic QM by a
2058: mathematical trick but is impossible in relativistic case. So what we
2059: need is a new understanding on the essence of special relativity
2060: --- the invariance (of theory) under the (newly defined) space-time
2061: inversion in one inertial frame. Then we are able to claim the same
2062: invariance in the RMCS with $\mu$ replacing $m$
2063: for establishing the RDE, ignoring a small centripetal acceleration of the nucleus in $CMCS$ (see page note after Eq.(\ref{3-22})).
2064: The approximation in $RDE$ is some price paid for the much
2065: bigger gain---improving the original Dirac equation (unable to treat
2066: different nuclei) and avoiding the confusion in QED calculation
2067: because of the entanglement of two frames: CMCS with RMCS (
2068: {\it i.e.}, the radiative corrections are entangled with the recoil effect
2069: as we can see from previous literatures). In some senses, we jump over
2070: obstacles (1) ,(2) and (4)at the least labor cost (by constructing
2071: RDE). In the meantime, we hope RDE would help to ease the
2072: difficulty in point (3).
2073: And it's a great pleasure to see that the essential correctness of our
2074: understanding has been validated by Marsch's work as well as
2075: puzzles raised in his paper \cite{34}. Please see also Ref \cite{41}.
2076: 
2077:    As to challenge (3), only after we puzzled over the "divergence"
2078: for decades, could we suddenly realize that we misread its implication
2079: as a "large number".
2080: Rather, it means the "uncertainty". Let us look at the calculation in
2081: section IV again. Previously, many authors treated the divergent
2082: integral I in Eq.(\ref{4-5}) by different tricks of regularization ,
2083: arriving at Eq.(\ref{4-11}). Because both A and B are divergent, it was
2084: thought that the original mass ($\mu$ here) does receive some radiative
2085: corrections (via the self-energy diagrams in Fig 2(a) and (b)) and
2086: becomes a "renormalized" mass ($\mu_R$ here). The latter should be the
2087: observed mass in experiment or physical mass (of electron). So the
2088: original mass was called as the "bare mass", which was written into
2089: the Lagrangian density as an input parameter of QFT. Then in
2090: constructing Feynman diagrams of certain perturbative calculation, one
2091: needs to further introduce some (divergent) "counter terms" for
2092: cancelling the divergence stemming from the bare mass. Based on that understanding, the renormalization
2093: factor for wavefunction, $Z_2$ in Eq.(\ref{4-10}), would be a divergent quantity
2094: too (in sharp contrast to here Eq.(\ref{4-14}) being a fixed
2095: and finite number). Previously, In Eq.(\ref{5-10}), while the $e_R$ on the left
2096: handed side is the observed charge which should be finite, the $e$ on
2097: the right handed side was regarded as a "bare charge" which, together
2098: with the $Z_3$, was a divergent quantity. (see Fig.~7.8 in \cite{23}. By contrast, here both $Z_3$ and
2099: $e$ are finite. Actually, here $e$ is defined as the physical charge
2100: observed at the Thomson limit in experiment).
2101: 
2102:   In our opinion, the reason why we encountered so many  superfluous troubles
2103: in the past is because we overlooked what Bethe said in 1947\cite{29}.
2104: Please read his words quoted after Eq. (A.2) in the Appendix. Let us
2105: explain via our Eq.(\ref{4-1}).
2106: The (reduced) mass ($\mu$) already contains some contributions from
2107: self-energy diagrams like Fig. 2(a) and (b). When we evaluate the
2108: (divergent) integral, Eq.(\ref{4-5}), trying to find the radiative
2109: corrections on the electron, the latter is bound to confuse with that
2110: already contained in the mass. In other words, the dividing line
2111: between them is blurred inevitably. The emergence of explicit
2112: divergence is essentially a warning that the effect you want to
2113: evaluate has been entangled with the mass itself, rendering both of
2114: them uncertain. Hence the aim of so-called renormalization is nothing
2115: but to redraw the dividing line between them such that the values of
2116: mass (reconfirmed by the experiment) and the new effect ({\it e.g.}, the
2117: mass increment when the electron is moving off-mass-shell, Eq.(\ref{4-16}))
2118: can be clearly separated.
2119:   In short, what we have been learning in the past decade is:
2120: At the level of QM, in the Hamiltonian like Eq. (A.1), the parameters
2121: $m$ and $e$ can be regarded as well-defined. But they are not so at the
2122: level of QFT. Once the calculation is made beyond the tree level,
2123: {\it i.e.}, with loop number $L\geq1$, the divergence occurs and the meaning of
2124: parameters becomes obscure immediately.
2125: 
2126: We need to reconfirm all parameters contained in the Lagrangian
2127: density before they can be linked with experiments. In this sense, a
2128: model of QFT is at most an "effective field theory". According to the
2129: above point of view, we believe that our RRM just provides a natural
2130: way to carry out these processes of reconfirmation \cite{8:a}, getting rid
2131: of divergences and ambiguities. Please see also Ref \cite{42}.
2132: 
2133: (h) Last, but not least, during the learning and teaching of graduate courses
2134: on QFT for decades, we have been sharing the joy and puzzle with
2135: our students all the time. We hope that the presentation of this
2136: paper could be useful as a teaching reference to render the QFT
2137: course more understandable, interesting and attractive.
2138: 
2139: \section*{Acknowledgements}
2140: 
2141: We thank S.Q. Chen, S.S. Feng, R.T. Fu,
2142: P.T. Leung, W.F. Lu, X.T. Song, F. Wang, H.B. Wang, J. Yan, G.H. Yang and
2143: J.F. Yang for close collaboration and helpful discussions.
2144: We are also indebted to referees whose comments provided us opportunities to
2145: improve our manuscript.
2146: 
2147: 
2148: \section*{Appendix: Comparison Between Noncovariant and Covariant Theories for Lamb Shift}
2149: 1. To our knowledge, the precision theory for Lamb shift was based on a combination of noncovariant
2150: (nonrelativistic or old-fashioned) QED with covariant (or relativistic) QED as discussed in Ref.\cite{27}.
2151: As explained clearly by Sakurai in Ref.\cite{21}, in perturbative QFT of noncovariant form, all virtual
2152: particles are "on-mass-shell". Here we wish to emphasize that a rigorous reconfirmation procedure of mass
2153: parameter was often overlooked in previous literatures.
2154: The theory for hydrogenlike atom begins with a Hamiltonian:
2155: \begin{equation*}\label{A1}
2156:     H_0=\frac{1}{2m}{\mathbf p}^2+\frac{1}{2m_N}{\mathbf p}^2-\frac{Z\alpha}{r}\eqno{(A.1)}
2157: \end{equation*}
2158: (${\mathbf p}=-i\nabla$, see Eq.(34) in \cite{27}). As Bethe \cite{29} first pointed out that the effect of
2159: electron's interaction with the vector potential $\mathbf A$ of radiation field (see \cite{21},\cite{15})
2160: \begin{equation*}\label{A2}
2161:     H_{int}^{(1)}=\frac{e}{mc}{\mathbf A}\cdot{\mathbf p}\eqno{(A.2)}
2162: \end{equation*}
2163: should properly be regarded as already included in the observed mass $m_{obs}$ of the electron, which is denoted
2164: by $m$ in (A1). However, once a concrete calculation is made with (A2) being taken into account, the
2165: divergence emerges immediately. What does it mean?
2166: Mathematicians teach us that there are three implications for a divergence:\\
2167: (a)It is a dimensionless number; (b)It is a large number; (c)It is uncertain. While we physicists often emphasized
2168: the point (b), we didn't pay enough attention to the points (a) and (c). We often talked about a quadratically
2169: (or linearly) divergent integral without noticing that it has a dimension (say, mass dimension) and thus
2170: meaningless in mathematics unless a mass parameter (say, $m$) in the integral is already fixed as a mass "unit"
2171: so that the integral can be divided by $m^2$ (or $m$) to become dimensionless.
2172: Alternatively, a logarithmically divergent integral is dimensionless and thus unaffected by the choice of unit
2173: [like Eq.(\ref{4-5}), see also Eq.(A6) below], it just implies an uncertainty waiting to be fixed.
2174: The implication of uncertainty of a divergence will never vanish even after we introduced a cutoff by hand to
2175: curb it. For example, in a pioneering paper to explain the Lamb shift, Welton (\cite{30}, see section 9.6B in
2176: Ref.\cite{15}) encountered an integral $I=\int_{\omega_{min}}^{\omega_{max}}\frac{d\omega}{\omega}$ with $\omega$
2177: being the (angular) frequency of virtual photon (vacuum fluctuation). He simply set
2178: $\omega_{min}\sim mZ\alpha =Z/a$, ($a$ is Bohr radius) and $\omega_{max}\sim m$ so that
2179: $I\simeq \ln(1/Z\alpha)=4.92$ (for $Z=1$) which leads to an estimation of Lamb shift
2180: $L_H^{theor.}(2S_{1/2}-2P_{1/2})\simeq 668\ MHz$. If instead of Bohr radius, the lower cutoff is provided by the
2181: electron binding energy, one should get $I\simeq \ln(Z\alpha)^{-2}$ and $L_H^{theor.}\simeq 1336\ MHz$.
2182: (see Eq.(30) in \cite{27}).
2183: The above arbitrariness just reflects what essential in a divergent integral is not its large magnitude
2184: ($\ln(Z\alpha)^{-1}$ is merely of the order of 10) but its uncertainty. So what important in handling the integral
2185: is not to curb (or to hide) its divergence but let the divergence exhibits itself as some arbitrary constants
2186: explicitly (as shown in section IV-VI). We will show later how to do this way for noncovariant QED.
2187: \\
2188: 2. While Eqs.(A1) and (A2) only describe a spinless particle, the electron has spin which endows it
2189: with the relativistic nature as shown by Eqs.(\ref{3-21})-(\ref{3-25}). For two-particle system,
2190: based on Bethe-Salpeter
2191: equation, an effective Dirac equation (EDE) was derived as shown by Eq.(23) in \cite{27}. When the electromagnetic
2192: field interaction is taken into account, the Breit potential $V_{Br}$ was derived as shown by Eq.{(35)}
2193: in \cite{27}.
2194: Then the total Breit Hamiltonian reads (Eq.(36) in \cite{27}):
2195: \begin{equation*}\label{A3}
2196:     H_{Br}=H_0+V_{Br}\eqno{(A.3)}
2197: \end{equation*}
2198: However, the electron mass $m'$ (in our notation here) appeared in EDE or $V_{Br}$ should be that in the Dirac
2199: equation, also that in the definition of reduced mass $\mu=\frac{m'm_N}{m'+m_N}$, eventually $m'$ could be
2200: identified with the observed mass $m_{obs}$, which is not equal to the $m$ in Eq.(A1). This is because besides
2201: (A2) there is an extra interaction due to electron spin with the radiation field:
2202: \begin{equation*}\label{A4}
2203:     H_{int}^{(2)}=\frac{ge\hbar}{4\mu c}{\mathbf \sigma}\cdot\nabla\times {\mathbf A}\eqno{(A.4)}
2204: \end{equation*}
2205: ($g=2\times 1.0011596522$ is the gyromagnetic ratio of electron, see Eq.(9A.15) in \cite{15}). The
2206: difference between
2207:  $m$ and $m'$ will be calculated in (A16) below. It turns out to be
2208: of the order of $\alpha m$ and cannot be ignored at the level of QED, especially for the explanation of
2209: Lamb shift.
2210: We guess this must be one of the reasons why all calculations based on Eq.(A3) became so complicated.
2211: \\
2212: 3. In noncovariant theory, the leading contribution to the Lamb shift comes from the one-photon electron
2213: self-energy.
2214: The nomenclature here is different from that in the covariant theory. Roughly speaking, so-called electron
2215: self-energy often corresponds to the vertex function in covariant theory (Fig.2(d) in this paper)
2216: or to Figs.8 and 11
2217: in Ref.\cite{27} and its evaluations have extended over 50 years \cite{31}. More precisely, it is identified with
2218: the radiative insertions in the electron line and the Dirac form factor contribution. Further contributions from
2219: the Pauli form factor and the vacuum polarization \cite{27} will add to a theoretical value of classic Lamb shift
2220: being $1050.559\ MHz$.
2221: If taking more high-order corrections into account, the theoretical value coincides with the experimental value
2222: $1057.845\ MHz$ rather accurately (see Table 20 in \cite{27}).
2223: However, the above calculation looks quite complicated due to two reasons: (a) The difficulty of dealing with
2224: two masses in two coordinate systems, the electron mass $m$ and the reduced mass $\mu$; (b) The introduction of
2225: an auxiliary parameter $\sigma\ [m(Z\alpha)\gg \sigma\gg m(Z\alpha)^2]$ to separate the radiative photon
2226: integration
2227: region into two parts. In the low momentum region, the Bethe Logarithm \cite{32} in noncovariant form makes the
2228: main contribution. In the high momentum region, the evaluation is resorting to some relativistic covariant form
2229: \cite{22}. Then two expressions are matched together to get the correct result.
2230: It seems to us that the matching trick used is doubtful because both ultraviolet and infrared divergences were
2231: ambiguously handled by some cutoff which missed the main point of renormalization---to reconfirm the mass parameter
2232: in the presence of radiative corrections as shown in section IV (covariant form) or below.
2233: \\
2234: 4. A simple calculation for Lamb shift in noncovariant form was proposed in Ref.\cite{26} (see also Appendix
2235:  9A of Ref.\cite{15}). Consider the self-energy diagram of an electron with reduced mass $\mu$ and
2236:  (three-dimensional)
2237:  momentum $\mathbf p$ in the RMCS of a hydrogenlike atom. Similar to Fig.2(a), but also different in the virtual
2238:  state, now a photon has energy $\omega_k=k=|{\mathbf k}|$ while the electron has momentum
2239:  ${\mathbf q}={\mathbf p}-{\mathbf k}$ and energy $\varepsilon_q=\frac{1}{2\mu}q^2$.
2240:  The electron in plane-wave state $|{\mathbf p}>$ has two interactions with the radiative field at each vertex as
2241:  shown by (A2) ($m\rightarrow \mu$) and (A4), acquiring an increase in energy respectively (see FIG. 3):
2242: \begin{equation*}\label{A5}
2243:     \Delta E_p^{(j)}=\sum_i\frac{|<i|H_{int}^{(j)}|{\mathbf p}>|^2}{\varepsilon_p-\varepsilon_i},
2244: \quad (j=1,2)\eqno{(A.5)}
2245: \end{equation*}
2246: Here $\varepsilon_i=\varepsilon_q+\omega_k$ is the energy of the intermediate virtual state $|i>$.
2247: Simple evaluation leads to
2248: \begin{equation*}\label{A6}
2249:     \Delta E_p^{(1)}=-\frac{\alpha p}{\pi\mu}\int_{-1}^1d\eta(1-\eta^2)I,
2250: \quad I=\int_0^{\infty}\frac{dk}{k+\xi}\eqno{(A.6)}
2251: \end{equation*}
2252: where $\eta=\cos\theta$ with $\theta$ being the angle between $\mathbf k$ and $\mathbf p$, $\xi=2(\mu-p\eta)$.
2253: Like Eq.(\ref{4-5}), we take partial derivative of the divergent integral $I$ with respect to $\xi$ (then the
2254: integration of $k$) and integrate back to $I$ again, yielding:
2255: $$
2256:   \Delta E_p^{(1)} = b_1^{(1)}p^2+b_2^{(1)}p^4+\cdots \eqno{(A.7)}
2257: $$
2258: 
2259:   $$b_1^{(1)} = \frac{\alpha}{\pi\mu}(\frac{4}{3}\ln2+\frac{4}{3}\ln\mu-\frac{4}{3}C_1) \eqno{(A.8)}$$
2260: 
2261:   $$b_2^{(1)} = \frac{\alpha}{\pi\mu^3}(-\frac{2}{15})\eqno{(A.9)}$$
2262: Note that the term $b_1^{(1)}p^2$ will combine with the kinetic energy $\frac{1}{2\mu}p^2$ of a ("spinless")
2263: electron, they are indistinguishable. The appearance of an arbitrary constant $C_1$ precisely reflects the fact
2264: that we cannot find the reduced mass via the valuation of $\Delta E_p^{(1)}$ in perturbation theory. So we must
2265: choose $b_1^{(1)}=0$ to reconfirm the value of $\mu$ (which is still not the final observed mass, see below).
2266: Similar evaluation on $H_{int}^{(2)}$ (of the real electron with spin $1/2$) which would induce the spin flip
2267: process between states $|{\mathbf p},\pm\frac{1}{2}>$ and $|{\mathbf q},\pm\frac{1}{2}>$, leads to
2268: $$
2269:   \Delta E_p^{(2)} =\frac{1}{2}\sum_{i,s_z=\pm1/2}\frac{|<i|H_{int}^{(2)}|{\mathbf p},s_z>|^2}
2270:   {\varepsilon_p-\varepsilon_i}=-\frac{\alpha g^2}{8\pi\mu}\int_{-1}^1d\eta J
2271: $$
2272: 
2273: $$  J =\int_0^{\infty}\frac{k^2dk}{k+\xi}\eqno{(A.10)}$$
2274: Being a quadratically divergent integral, $J$ needs partial derivative of third order with respect to $\xi$,
2275: yielding:
2276: $$
2277:  \Delta E_p^{(2)} =b_0^{(2)}+b_1^{(2)}p^2+b_2^{(2)}p^4+\cdots \eqno{(A.11)} $$
2278: 
2279:  $$
2280:  b_0^{(2)} = \frac{g^2}{4}\frac{\alpha\mu}{\pi}[4(\ln2+\ln\mu)-4C_2-\frac{2C_3}{\mu}-\frac{C_4}{\mu^2}]\eqno{(A.12)}
2281:  $$
2282: 
2283:  $$
2284:   b_1^{(2)} =\frac{g^2}{4}\frac{\alpha}{\pi\mu}(\frac{4}{3}\ln2+2+\frac{4}{3}\ln\mu-\frac{4}{3}C_2) \eqno{(A.13)}
2285:   $$
2286: 
2287:  $$ b_2^{(2)} = \frac{g^2}{4}\frac{\alpha}{\pi\mu^3}(-\frac{1}{15})\eqno{(A.14)}$$
2288: Let's manage to fix three arbitrary constants $C_2,C_3$ and $C_4$. First, the term $b_1^{(2)}p^2$ should be
2289: combined
2290: with $\frac{1}{2\mu}p^2$ term. Since $\mu$ is already fixed, further modification on $\mu$ due to electron spin
2291: should be finite and fixed. So the only possible choice of $C_2$ is to cancel $\ln\mu$ which is ambiguous in
2292: dimension: $C_2=\ln\mu$, yielding
2293: \begin{equation*}\label{A15}
2294:     b_1^{(2)}=\frac{\beta}{2\mu},\quad \beta=\frac{g^2\alpha}{2\pi}(\frac{4}{3}\ln2+2)\eqno{(A.15)}
2295: \end{equation*}
2296: Then the dimensional constants $C_3$ and $C_4$ must be chosen such that $b_0^{(2)}=0$, implying that the starting
2297: point of this theory is the nonrelativistic Hamiltonian $H_0$ in Eq.(A1) without rest energy term while both
2298: masses of the nucleus and the electron (with spin) are fixed by experiments.
2299: Hence now $\mu$ acquires a modification via $b_1^{(2)}p^2$ term and becomes an observable one:
2300: \begin{equation*}\label{A16}
2301:    \mu\longrightarrow \mu_{obs}=\frac{\mu}{1+\beta}\eqno{(A.16)}
2302: \end{equation*}
2303: However, we have to consider the relativistic energy of electron shown in Eq.(\ref{3-10}), where the term
2304: $(-\frac{1}{8\mu^3}p^4)$ goes beyond Eq.(A1). Yet the modification of $\mu$ shown as (A16) does induce
2305: a corresponding change $-\frac{1}{8}(\frac{1}{\mu_{obs}^3}-\frac{1}{\mu^3})p^4$, which should be regarded as
2306: an invisible "background" and subtracted from the $p^4$ term induced by radiative corrections. (The relativistic
2307: correction is brought in via the RDE as discussed in section VIII). As a whole, the combination of contributions
2308: from $H_{int}^{(1)}$ and $H_{int}^{(2)}$ leads to
2309: \begin{equation*}\label{A17}
2310:    b_1= b_1^{(1)}+ b_1^{(2)}= b_1^{(2)}\eqno{(A.17)}
2311: \end{equation*}
2312: and a "renormalized" $b_2$:
2313: \begin{equation*}\label{A18}
2314:    b_2^R=b_2^{(1)}+b_2^{(2)}+\frac{1}{8\mu^3}(3\beta+3\beta^2+\beta^3)
2315:    \simeq\frac{\alpha}{\pi\mu_{obs}^3}(1.99808)\eqno{(A.18)}
2316: \end{equation*}
2317: Here we only keep the lowest approximation at the last step.
2318: Hence the electron self-energy-diagram contributes a radiative correction to the energy level of the
2319: stationary state $|Z,n,l>$ in a hydrogenlike atom:
2320: \begin{equation*}\label{A19}
2321:    \Delta E^{rad}(Z,n,l)=\langle Z,n,l|b_2^Rp^4|Z,n,l\rangle=[\frac{8n}{2l+1}-3]\frac{b_2^RZ^4\alpha^4}{n^4}\mu^4_{obs}\eqno{(A.19)}
2322: \end{equation*}
2323: This form, together with contributions from the vacuum polarization and nuclear size effect, gives a theoretical
2324: value for classic Lamb shift:
2325: \begin{equation*}\label{A20}
2326:    L_H^{theor.}(2S_{1/2}-2P_{1/2})\approx 1056.52\ MHz\eqno{(A.20)}
2327: \end{equation*}
2328: which is smaller than the experimental value by $0.13\%$.
2329: Despite its approximation involved, the above method clearly shows that so-called renormalization is nothing but a
2330: reconfirmation process of mass. We must reconfirm the mass before it could be modified via radiative corrections.
2331: Either "skipping over the first step" or "combining two steps into one" is not allowed.
2332: \\
2333: 5. In noncovariant theory, the (three-dimensional) momentum $\mathbf p$ is combined with the reduced mass $\mu$
2334: to form a kinetic energy term $\frac{1}{2\mu}{\mathbf p}^2$ on the mass shell. Once the energy is modified
2335: whereas $\mathbf p$ is conserved at the vertex, $\mu$ is bound to be modified. On the other hand, in covariant
2336: theory, the electron energy turns to a component of four-dimensional momentum $p$ and the latter is conserved
2337: at the vertex. So the (reduced) mass  $\mu$  cannot be modified on the mass shell $(p^2=\mu^2)$. Therefore,
2338: the renormalization as some reconfirmation has different meaning in covariant theory versus that in
2339: noncovariant theory. We guess this is why the matching procedure of these two formalisms into one theory
2340: for Lamb shift proves so difficult.
2341: \\
2342: 6. Every theory in physics is not only a discovery of natural law, but also an invention of human being \cite{33}.
2343: Hence the comparison among various theories, in many cases, is not about a problem of being right or wrong.
2344: Rather, it's about a choice of simplicity, harmony (self-consistency) and beauty. Only time can tell.
2345: 
2346: 
2347: 
2348: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2349: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2350: %                      BIBLIOGRAPHY
2351: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2352: \begin{thebibliography}{99}
2353: 
2354: \bibitem{1}
2355: Sapirstein J R and Yennie D R 1990 in {\it Quantum
2356: Electrodynamics} (Editor: Kinoshita T, World Scientific )
2357: p.560.
2358: 
2359: \bibitem{2}
2360: Pachucki K \etal\ \jpb{29}{1996}{177}.
2361: 
2362: \bibitem{3}
2363: Barker W A  and Chraplyvy Z V \pr{89}{1953}{446};\\
2364: Chraplyvy Z V \pr{91}{1953}{388}; \pr{92}{1953}{1310}.
2365: 
2366: \bibitem{4}
2367: Barker W A and Glover F N \pr{99}{1955}{317}.
2368: 
2369: \bibitem{5}
2370: Udem Th \etal\ \prl{79}{1997}{2646}.
2371: 
2372: \bibitem{6}
2373: Schmidt-Kaler F \etal\ \prl{70}{1993}{2261}.
2374: 
2375: \bibitem{7}
2376: Yang J F 1994 Thesis for PhD (Fudan University); hep-th/9708104;\\
2377: Yang J F and Ni G J 1995 {\it Acta Physica Sinica} (Overseas Edition) {\bf
2378: 4} 88.
2379: 
2380: \bibitem[8a]{8:a}
2381: Ni G J and Chen S Q 1998 {\it Acta Physica Sinica} (Overseas Edition) {\bf
2382: 7} 401.
2383: 
2384: \bibitem[8b]{8:b}
2385: Ni G J, Yang G H, Fu R T and Wang H B 2001 {\it Int. J. Mod. Phys.
2386: A} {\bf 16} 2873; \hepth{9906254}.
2387: 
2388: \bibitem[9]{9}
2389: Yang J F \prd{57}{1998}{4686}.
2390: 
2391: \bibitem[10]{10}
2392: Weitz M  \etal\ \pra{52}{1995}{2664}.
2393: 
2394: \bibitem[11]{11}
2395: Kinoshita T \prl{75}{1995}{4728}.
2396: 
2397: \bibitem[12]{12}
2398: Farnham D L \etal\ \prl{75}{1995}{3598}.
2399: 
2400: \bibitem[13]{13}
2401: Review of Particle Physics 2000 {\it Euro. Phys. J. C} {\bf 15} 350.
2402: 
2403: \bibitem[14]{14}
2404: Ni G J 2003 {\it Progress in Physics} (Nanjing, China) {\bf 23} 484
2405: ; \hepth{0206250}.
2406: 
2407: \bibitem[15]{15}
2408: Ni G J and Chen S Q 2003 {\it Advanced Quantum Mechanics 2nd.} (Fudan University Press); English Edition was
2409: published by the Rinton Press, 2002.
2410: 
2411: \bibitem[16]{16}
2412: Ni G J 2003 in {\it Relativity, Gravitation, Cosmology}, Edit by Dvoeglazov V V and Espinoza Garrido A A,
2413: Chapter 10 (NOVA Science Publisher); \physics{0302038}.
2414: 
2415: \bibitem[17]{17}
2416: Ni G J 2003 in {\it Relativity, Gravitation, Cosmology}, Edit by Dvoeglazov V V and Espinoza Garrido A A,
2417: Chapter 11 (NOVA Science Publisher); \hepph{0306028}; preprint, \hepph{0404030}, to be published
2418: in {\it Relativity, Gravitation, Cosmology}.
2419: 
2420: \bibitem[18]{18}
2421: Ni G J and Chen S Q 1997 {\it J. Fudan University} (Natural Science)
2422: {\bf 36} 247; \hepth{9708156};\\
2423: in {\it Proceedings of International Workshop, Lorentz Group, CPT and
2424: Neutrinos}, Zacatocas, Maxico, June 23-26, 1999 (World Scientific,
2425: 2000), p.450.
2426: 
2427: \bibitem[19]{19}
2428: Gross D J and Neveu A \prd{10}{1974}{3235}.
2429: 
2430: \bibitem[20]{20}
2431: Ni G J and Wang H B 1998 {\it Physics Since Parity Breaking}, edited by
2432: Wang F (World Scientific, Singapore) p.436; \hepph{9708457}.
2433: 
2434: \bibitem[21]{21}
2435: Sakurai J J 1967 {\it Advanced Quantum Mechanics} (Addison-Wesley Publishing Company).
2436: 
2437: \bibitem[22]{22}
2438: Itzykson C  and Zuber J-B 1980 {\it Quantum Field Theory} (McGraw-Hill Book Company).
2439: 
2440: \bibitem[23]{23}
2441: Peskin M E  and Schroeder D V 1995 {\it An Introduction to Quantum Field Theory} (Addison-Wesley Publishing Company).
2442: 
2443: \bibitem[24]{24}
2444: Ni G J, Lou S Y, Lu W F and Yang J F 1998 {\it Science in China}
2445: (Series A), {\bf 41} 1206; \hepph{9801264}.
2446: 
2447: \bibitem[25]{25}
2448: Feng S S  and Ni G J 1999 {\it Int. J. Mod. Phys. A} {\bf 14} 4259.
2449: 
2450: \bibitem[26]{26}
2451: Ni G J, Wang H B, Yan J and Li H L 2000 {\it High Energy Physics and
2452: Nuclear Physics}{\bf 24} 400.
2453: 
2454: \bibitem[27]{27}
2455: Eides M I, Grotch H and Shelyuto V A \prep{342}{2001}{63-261}.
2456: 
2457: \bibitem[28]{28}
2458: Huber A , Udem Th, Gross B  \etal\ \prl{80}{1998}{468}.
2459: 
2460: \bibitem[29]{29}
2461: Bethe H A \pr{72}{1947}{339}.
2462: 
2463: \bibitem[30]{30}
2464: T. A. Welton \pr{74}{1948}{1157}.
2465: 
2466: \bibitem[31]{31}
2467: Jentschura U D , Mohr P  J  and Soff G \prl{82}{1999}{53}.
2468: 
2469: \bibitem[32]{32}
2470: Drake G  W  F   and Swainson R  A \pra{41}{1990}{1243}.
2471: 
2472: \bibitem[33]{33}
2473: Ni G J 2004 \physics{0407092} accepted by "{\it Relativity, Gravitation, Cosmology}".
2474: 
2475: \bibitem[34]{34}
2476: Marsch E 2005 {\it Ann. Phys.} (Leipzig) {\bf 14} No.5 324.
2477: 
2478: \bibitem[35]{35}
2479: Dai Xianxi 1977 {\it Journal of Fudan University} No.1 100.
2480: 
2481: \bibitem[36]{36}
2482: Ni G J and Chen S Q 1995 {\it Levinson Theorem, Anomaly and the Phase Transition of Vacuum}
2483: (Shanghai Scientific \& Technical Publishers).
2484: 
2485: \bibitem[37]{37}
2486:  Ma Z Q  and  Ni G J \prd{31}{1985}{1482}.
2487: 
2488: \bibitem[38]{38}
2489: Weinberg S 1995 {\it The Quantum Theory of Fields} Vol.I-II (Cambridge University Press, Cambridge)
2490: 
2491: \bibitem[39]{39}
2492: Amsler C \etal\ {\it Particle Physics Booklet} \plb{667}{2008}{1}
2493: 
2494: \bibitem[40]{40}
2495: Lou S Y  and Ni G J \prd{40}{1989}{3040}
2496: 
2497: \bibitem[41]{41}
2498: Ni G J, Chen S Q, Lou S Y and Xu J J  {\it Essence of Special
2499: Relativity, Reduced Dirac Equation and Antigravity},\ Preprint.
2500: 
2501: \bibitem[42]{42} Ni G J, Xu J J and Lou S Y {\it Puzzles of Divergence and
2502: Renormalization in Quantum Field Theory},\ Preprint.
2503: 
2504: \bibitem[43]{43} 
2505: Niering M \etal,  \prl{84}{2000}{5496}
2506: 
2507: \bibitem[44]{44} 
2508: Schwob C \etal,  \prl{82}{1999}{4960}
2509: 
2510: \bibitem[45]{45} 
2511: de Beauvoir B \etal, 2000 {\it Eur. Phys. J} D {\bf 12} 61
2512: 
2513: 
2514: \end{thebibliography}
2515: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2517: \begin{figure*}[!h]
2518: \centerline{\includegraphics[scale=0.8]{fig1.pdf}}
2519: \vskip 0.5cm
2520: \caption{
2521: A hydrogenlike atom in quantum mechanical description. The nucleus with mass $m_2$ occupies a small sphere with radius $r_N$
2522: (greatly exaggerated in the diagram) while the electron with mass $m_1$ spreads over a larger sphere with radius $R_e$
2523: (\ie atomic radius). Their common center is the atom's center of mass (CM). The wavefunction $\psi({\bf r})e^{-iEt}$ with
2524: ${\bf r}={\bf r}_1-{\bf r}_2$ shows the electron's amplitude under a "fictitious measurement" \cite{15}, during which the
2525: electron and nucleus shrink into two "fictitious point particles " located at ${\bf r}_1$ and ${\bf r}_2$ simultaneously.
2526: The Coulomb potential $V(r)=-\frac{Ze^2}{r}$ between them is a static one. The probability
2527: to find the electron at ${\bf r}$ is $|\psi({\bf r})|^2$ while that to find its momentum being ${\bf p}$ is $|\phi({\bf p})|^2$ with
2528: $\phi({\bf p})$ being the Fourier transform of $\psi({\bf r})$.
2529: }
2530: \end{figure*}
2531: 
2532: \begin{figure*}[!h]
2533: \centerline{\includegraphics[scale=0.8]{fig2.pdf}}
2534: \vskip 0.5cm
2535: \caption{Four Feynman diagrams at one-loop level (in covariant form). (a) and (b) are self-energy diagrams of the electron.
2536: (c) is vacuum polarization. (d) is vertex function. Solid lines and wavy lines refer to electron and photon
2537: respectively, while X denotes the nucleus. Here $p,q$ and $k$ are
2538: four-dimensional momenta.}
2539: \end{figure*}
2540: 
2541: \begin{figure*}[!h]
2542: \centerline{\includegraphics[scale=1.3]{e1.pdf}}
2543: \vskip 0.5cm
2544: \caption{
2545: The electron self-energy (radiative correction) diagram at one-loop level of perturbative QCD in noncovariant form.
2546: $H_{int}^{(1)}$ (A.2) or $H_{int}^{(2)}$ (A.4) is inserted into two vertices. Here $\bf p,q$ and $\bf k$ are three
2547: dimensional momenta.
2548: }
2549: \end{figure*}
2550: 
2551: %\end{CJK*}
2552: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2553: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2554: \end{document}
2555: